paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
1611.08751
1
1611
2016-11-26T22:24:29
Functional Alignment with Anatomical Networks is Associated with Cognitive Flexibility
[ "q-bio.NC" ]
Cognitive flexibility describes the human ability to switch between modes of mental function to achieve goals. Mental switching is accompanied by transient changes in brain activity, which must occur atop an anatomical architecture that bridges disparate cortical and subcortical regions by underlying white matter tracts. However, an integrated perspective regarding how white matter networks might constrain brain dynamics during cognitive processes requiring flexibility has remained elusive. To address this challenge, we applied emerging tools from graph signal processing to decompose BOLD signals based on diffusion imaging tractography in 28 individuals performing a perceptual task that probed cognitive flexibility. We found that the alignment between functional signals and the architecture of the underlying white matter network was associated with greater cognitive flexibility across subjects. Signals with behaviorally-relevant alignment were concentrated in the basal ganglia and anterior cingulate cortex, consistent with cortico-striatal mechanisms of cognitive flexibility. Importantly, these findings are not accessible to unimodal analyses of functional or anatomical neuroimaging alone. Instead, by taking a generalizable and concise reduction of multimodal neuroimaging data, we uncover an integrated structure-function driver of human behavior.
q-bio.NC
q-bio
Functional Alignment with Anatomical Networks is Associated with Cognitive Flexibility John D. Medaglia1, Weiyu Huang2, Elisabeth A. Karuza1, Sharon L. Thompson-Schill1, Alejandro Ribeiro2, and Danielle S. Bassett2,3 1Department of Psychology, University of Pennsylvania, Philadelphia, PA, 19104 USA 2Department of Electrical & Systems Engineering, University of Pennsylvania, Philadelphia, PA, 3Department of Bioengineering, University of Pennsylvania, Philadelphia, PA, 19104 USA 19104 USA 6 1 0 2 v o N 6 2 ] . C N o i b - q [ 1 v 1 5 7 8 0 . 1 1 6 1 : v i X r a 1 Article summary Human cognitive flexibility emerges from complex dynamics in anatomical networks. How- ever, no concise measure integrating anatomical and functional organization in support of cognitive flexibility has been identified. We examine cognitive "switch costs" – a measure of cognitive flexibility – in 28 healthy individuals during functional neuroimaging. We con- struct individualized anatomical networks from diffusion spectrum imaging data, allowing us to reconstruct the white matter networks for each subject. We apply a novel approach from graph signal processing to distill functional brain signals into aligned and liberal components relative to underlying anatomy. We show that relatively alignment of brain signals with underlying anatomical network organization supports faster cognitive switching. Moreover, we show that the signals most aligned with the network from moment to moment are also the most flexible. This indicates that anatomical networks provide organization to the flexible brain and facilitate greater mental agility and establishes a potential neural biomarker that integrates brain structure and function. 2 Abstract Cognitive flexibility describes the human ability to switch between modes of mental function to achieve goals. Mental switching is accompanied by transient changes in brain activ- ity, which must occur atop an anatomical architecture that bridges disparate cortical and subcortical regions by underlying white matter tracts. However, an integrated perspec- tive regarding how white matter networks might constrain brain dynamics during cognitive processes requiring flexibility has remained elusive. To address this challenge, we applied emerging tools from graph signal processing to decompose BOLD signals based on diffusion imaging tractography in 28 individuals performing a perceptual task that probed cognitive flexibility. We found that the alignment between functional signals and the architecture of the underlying white matter network was associated with greater cognitive flexibility across subjects. Signals with behaviorally-relevant alignment were concentrated in the basal gan- glia and anterior cingulate cortex, consistent with cortico-striatal mechanisms of cognitive flexibility. Importantly, these findings are not accessible to unimodal analyses of functional or anatomical neuroimaging alone. Instead, by taking a generalizable and concise reduction of multimodal neuroimaging data, we uncover an integrated structure-function driver of hu- man behavior. 3 Introduction Cognitive flexibility is involved in virtually every complex behavior from mental arithmetic to processing visual stimuli. For example, when navigating complex environments, humans can flexibly switch between two foci of attention or between two processing modalities, in order to effectively respond to sensory inputs. While a hallmark of human cognition, flexible switching is also associated with a measurable cost: moving from one task to another induces a natural extension in the time it takes a person to respond [1]. In patients with neurological syndromes, this cost is even greater, to the point where it can hamper a patient's ability to engage in the basic activities of daily living [2], impacting long-term cognitive outcomes [3]. In healthy individuals, cognitive flexibility varies considerably, and individual differences in this trait contribute to mental facets ranging from the development of reasoning ability [4] to quality of life into late age [5]. Clarifying the nature of cognitive flexibility in the human brain is critical to understand the human mind. The physiological origins of cognitive flexibility are thought to lie in corticobasal ganglia- thalamo-cortical loops [6]: regions of the fronto-parietal and cingulo-opercular systems are activated by cognitive switching tasks [7, 8, 9, 10]. In switching paradigms, the anterior cin- gulate is thought to contribute negative feedback detection following switches, whereas the lateral prefrontal cortex maintains rules and inhibits incorrect responses. Both of these re- gions anatomically connect to subcortical regions, which are postulated to mediate processes that both suppress prepotent motor responses and transition between behavioral outputs to meet task goals [11]. Interactions between cortical systems and motor outputs are thought to be anatomically mediated by subcortical circuits [7, 12, 13, 14, 9]. Yet, understanding ex- actly how this circuit supports task switching has remained challenging, particularly because it requires a conceptual integration of regional activity, inter-regional anatomical connectiv- ity, and observable measures of behavior. While regional activity and behavioral markers of cognitive flexibility are relatively straightforward to estimate, it is less straightforward to integrate these features with the white matter structure (the connectome [15]) that guides the propagation of functional signals [16, 17, 18]. As a result, these two research enterprises – cognitive neuroscience and connectomics – have largely developed in parallel without significant cross-talk. Ideally, frameworks that include a concise correspondence between brain network structure, function, and cognitive measures have the potential to produce more comprehensive understanding of human cog- nition [19, 20]. Conceptually, underlying white matter network organization in the brain 4 physically mediates communication among brain regions. However, analytic frameworks that explicitly use white matter structure to constrain cognitively relevant functional signals are lacking. Such approaches may allow investigators to adjudicate the relative contributions of well-described systems in the brain [21, 8] to specific cognitive variables by integrating neurophysiological dynamics and anatomy. To address this challenge, we aim to identify the multimodal integration of network anatomy and functional signals that supports cognitive switching. We introduce a novel approach that allows us to examine the distinct contributions of functional signals in the context of anatomically linked regions in human brain networks. In a cohort of 28 healthy adult human subjects, we collected high-resolution diffusion spectrum imaging (DSI) data as well as BOLD fMRI data acquired during the performance of a novel cognitive switching paradigm built on a set of shapes that could be perceived as composed of different features at the local versus global scales [22] (see Fig. 1). From the DSI data, we constructed anatom- ical brain networks in which 111 cortical, subcortical, and cerebellar regions [23, 24] were linked pairwise by the density of streamlines reconstructed by state-of-the-art tractography algorithms. Next, we used the eigenspectrum of these anatomical networks to measure the relative separation of framewise regional BOLD signals from the underlying white matter (see Fig. 2 and Methods). Specifically, each regional signal was decomposed into a portion that aligned tightly with the anatomical network ("aligned") and a portion that did not align tightly with the network ("liberal"). Here, alignment and liberality refer to signal deviations from the underlying anatomical network. We examine the observed BOLD activity for each measurement in time to identify where and to what extent BOLD signals across the brain are organized by white matter networks. Conceptually, this technique allows us to identify to what degree individual BOLD signals deviate weakly versus strongly from underlying white matter anatomy. We anticipate that functional alignment with anatomical networks is an individually vari- able feature that facilitates cognitive flexibility. We hypothesize that moment-to-moment alignment in human brain networks facilitates switching performance, indicating interindi- vidual variability in the degree of organization of information processing by anatomical topology. This allows us to investigate a theoretical dynamic property supporting cognitive flexibility and provides a mapping to neuroanatomical theories of cognitive switching. In the current study, our task involves proactive switching following detection of a contextual cue [11]. Whereas prior literature has focused on region-specific mechanisms associated with this 5 process, the current approach allows us to examine the role of local neural processes across the brain's distributed anatomical network. We postulate that functional alignment serves as a mechanism for cognitive flexibility across the brain's complex anatomical white matter network. In particular, well-aligned signals may represent efficiently organized signals with markedly less interference, whereas liberal signals do not leverage the inherent organization of underlying white matter networks. This approach to multimodal neuroimaging could dis- tinguish whether structure and function operate synergistically or in opposition to promote cognitive flexibility. Accordingly, we examine whether alignment with anatomically medi- ated expectations in subcortical regions, cortical systems, or both constitutes the basis for effective cognitive switching. This allows us to determine (i) whether alignment forms a basis for cognitive flexibility and (ii) whether regions such as those in the subcortical system that serve an anatomically central network role also serve a greater or less functionally central role in this mental skill. Figure 1: Cognitive task requiring perceptual switching. (A) Example stimuli based on Navon local-global features. Subjects were trained to respond to the larger (or "global") shape if the stimulus was green and to the smaller (or "local") shapes if it was white. (B) An example of the non-switching condition for responses. Subjects viewed a sequence of images and were instructed to respond as quickly and accurately as possible. (C) An example of the switching condition between stimuli requiring global and local responses. Here, trials with a red exclamation point are switches from the previous stimulus. 6 Figure 2: Multimodal approach to the study of cognitive switching using emerging graph signal processing tools. (A) A notion of signal independence on a schematic modular network. Left: An aligned signal on top of a given graph is one in which the magnitude of functional signals, represented by the directionality of the colored cones, corresponds tightly to that expected by the network's organization. In this toy example, one cluster of nodes contains similar positive signals, and the other cluster contains similar negative signals. Right: A liberal signal on top of a given graph is one in which signals diverge significantly from the underlying network. (B) We construct a white matter graph (adjacency matrix) from 111 anatomically-defined regions where connections are the streamline density between region pairs. (C) From BOLD fMRI data acquired during the performance of the Navon task, we extract regional mean time series which we treat as graph signals. (D) For each subject, we decompose the graph signals into aligned and liberal components using the underlying eigenspectrum of the white matter graph. We spatially map aligned and liberal signals, and we determine their relationship to switch costs estimated from behavioral performance on the task. Cb = cerebellum. See Methods for details. Functional alignment in liberal subcortical systems is associated with faster cognitive switching Aligned signals were concentrated within default mode, fronto-parietal, cingulo-opercular, and subcortical systems across subjects, whereas the liberal signals were concentrated largely in the subcortical system (Fig. 4). The significance of these concentrations was confirmed statistically using a non-parametric permutation test (α = 0.05) in which we shuffled the 7 values of alignment (or liberality) uniformly at random across brain regions before computing the mean alignment (or liberality) value within each system [25]. Then, we calculated the correlation between aligned and liberal BOLD signals across the brain and cognitive switch costs (response times during trials with a color-cued switch versus non-switching trials). We observed that variability in aligned signals was not associated with switch costs (R = 0.15, p = 0.43, accounting for 2% of the variance), while variability in liberal signals accounted for 32% of variance (see Fig. 3). Among the liberal signals, lower values of liberality (that is, relative alignment) were also associated with lower switch costs both during fixation (R = 0.62, p = 0.0006) and during nonswitching (R = 0.71, p = 0.0001) perceptual blocks. Figure 3: Lower independence is associated with lower switch costs. (A) Liberal signals were concentrated especially in subcortical regions and cingulate cortices. (B) Reduced liberality (in- creased alignment) is associated with reduced switch costs across subjects. (C) Aligned signals were concentrated especially in subcortical, default mode, fronto-parietal, and cingulo-opercular systems. (D) Variability in aligned signals was not significantly associated with switch costs across subjects. L = left hemisphere, R = right hemisphere. It is interesting to note that signals extracted from subcortical areas can be decomposed into both highly anatomically-aligned portions and highly liberal portions (see Methods). The values of both alignment and liberality were significantly greater than expected in sub- cortical structures. Yet, other systems show a disposition toward high alignment (fronto- parietal, cingulo-opercular, and default mode systems), or no disposition toward alignment or liberality (visual, auditory, and dorsal & ventral attention). We quantified the relationship between aligned and liberal signals by calculating the Pearson correlation coefficient between alignment and liberality values over all brain regions. We observed that the two types of 8 Figure 4: Non-parametric permutation test for signal concentration within cognitive systems. (A) Liberal signal concentrations sorted from highest ( top) to lowest concentration across all regions in all subjects. (B) Liberal signals were most concentrated in subcortical regions. (C) Aligned signal concentrations sorted from highest ( top) to lowest concentration across all regions in all subjects. (D) Aligned signals were most concentrated in fronto-parietal, cingulo-opercular, default mode, and subcortical systems. 9 signals were significantly but not perfectly correlated across the brain: mean over subjects and tasks R = 0.69, p (cid:28) 0.001 (see also Supplement). This analysis highlights an interesting measure of multimodal complexity that differs across brain areas: subcortical regions can be thought of as displaying high multimodal complexity because they contain both highly anatomically-aligned signals and highly liberal signals. Anatomy and function together define a neural trait un- derlying cognitive flexibility Finally, we asked whether the observed regional variation in aligned and liberal signals could be estimated reliably across subjects and tasks, supporting its utility as a marker for behavior. To address this question, we compared the regional patterns of alignment between subjects and between task conditions (fixation, non-switching, and switching). We observed that the regional pattern of signal alignment was consistent across the three task conditions for a given subject (intraclass correlation: R = 0.99, p (cid:28) 0.001), as was the regional pattern of liberal signals (intraclass correlation: R = 0.99, p (cid:28) 0.001; see also Supplement). These results indicate that regional patterns of signal alignment and liberality are highly stable phenotypes of a subject across the three task conditions. Taken together, our results demonstrate that individuals whose most liberal functional signals were more aligned with white matter architecture could switch perceptual focus faster. In other words, relative alignment with anatomy was associated with greater cognitive flexi- bility, a finding that highlights the importance of simultaneously considering both functional and anatomical neuroimaging in the study of higher order cognitive processes. These findings complement prior studies of executive function that have focused on node-level, edge-level, and module-level features of brain networks [26, 13] (see also Supplement). Here, we consider brain function as a series of time-evolving states [25, 27] that are organized in relation to the underlying pattern of white matter tracts. The state-based focus of our approach also offers insights into the differential extent to which specific cognitive systems deviate from tract anatomy, underscoring anatomical contributions to the organization of brain dynamics across subjects. Our results contextualize previous models of cortico-striatal cognitive switching mecha- nisms [11, 28] within a connectomic perspective. As a complement to prior findings implicat- ing individual prefrontal and striatal systems in cognitive switching, our results highlight the 10 importance of anatomical network organization, and the central role of subcortical functional dynamics atop that structure. Specifically, we found that modest functional alignment of the most liberal signals in the subcortical and cingulate cortices was associated with faster cognitive switching. This observation is particularly interesting in the context of prior work showing that subcortical and anterior cingulate regions manage multiple inputs and outputs among cortical systems during task transitions [29, 9], potentially requiring more diverse signal organization relative to anatomical networks. Interestingly, we observe that the levels of alignment and liberality were consistent within subjects across all three conditions of the cognitive task – fixation, non-switching, and switching blocks – suggesting that it may be a trait rather than a state of an individual. While prior studies have similarly identified consistent network architecture across fixation and task [30], these similarities have not pre- viously been identified in a model-based integration of functional dynamics and white matter anatomy. The observed inter-task consistency of functional signal alignment and liberality was complemented by broad inter-subject variability. We speculate that this variability rep- resents differences in the amount of energy utilized to enact perceptual switches [31], with greater independence requiring greater energy. Thus, future studies could examine the po- tentially subject-specific role of cognitive control systems in regulating energy-efficient state transitions in local-global processing. With respect to recent dynamic network analyses of executive function, this analysis contributes a crucial anatomically-grounded perspective. The current approach represents a framework in which to understand the dual features of anatomical organization and func- tional processes supporting cognitive flexibility in the human connectome. Here, high func- tional dependence in fronto-parietal, cingulo-opercular, default mode, and subcortical sys- tems is not associated with intersubject switching variability. Critically, our results indicate that regions that participate in highly flexible systems [32] in temporal network analysis demonstrate high dependence on underlying anatomical networks across frames of BOLD data during fixation, low cognitive control conditions, and task conditions (see Supplement for further analysis and discussion). Previous studies identify dynamic network roles for fronto-parietal and cingulo-opercular regions in cognitive switching, and our results indicate that moment to moment signal configurations in highly flexible systems are strongly orga- nized by structure across time (see Supplement for addition analysis and discussion). In the context of this highly organized cortical activity, the current results suggest that subcortical systems contain highly liberal signals. The extent to which subcortical systems exhibit rela- 11 tive alignment may form a flexible integrative role across the many computations supported by cortical systems. The relationship between anatomically-bound momentary signal orga- nization and functional reconfigurations in temporal networks may more generally provide a fruitful area for future research. Notably, our results do not explain the potential cognitive role of highly aligned signals. It is possible that the role of these signals may be better explained in the context of other cognitive control processes [33]. We speculate that anatomically aligned signals in fronto- parietal, cingulo-opercular, default mode, and subcortical systems organize the dynamic signals contributing to cortical mechanisms of cognitive control, attention, and resting and preparedness processes, respectively. It would be interesting to test whether highly aligned signals in association cortices and subcortical structures are associated with domain-general performance variability across modalities [34]. In cognitive switching specifically, the extent to which signal liberality relates to performance on tasks involving other sensory modalities, transitioning between internally and externally focused attention, and divergent thinking remains to be established. In conclusion, our study supports the utility of network science in clarifying mechanisms of executive function specifically and cognition more generally [19, 20]. Recent literature firmly establishes that white matter organization is a critical, but incomplete determinant of functional signals in brain networks [17, 35]. Conceptually, the current approach acknowl- edges that without structure, functional signals lack a mediating organization. By examining functional signal alignment within underlying white matter networks, we identify an impor- tant definition of dynamic contributions to cognitive switching that powerfully discriminates between the contributions of subcortical and other systems in the brain. Similar applications to other large multimodal neuroimaging datasets could contribute to biomarker analyses in psychiatric disease and neurological disorders, many of which are associated with deficits in executive function [36, 37, 38]. 12 Methods Subjects A total of 30 subjects were recruited. All subjects were screened for prior history of psychi- atric or neurological illness. One subject was excluded due to near-chance performance on the task (accuracy = 52%). One additional subject was excluded due to technical problems on the day of scanning. The final sample included 28 individuals (mean age = 25.6, St.D. = 3.5, 70% caucasian, 13 females). All subjects volunteered with informed consent in writing in accordance with the Institutional Review Board/Human Subjects Committee, University of Pennsylvania. Behavioral task All participants completed a local-global perception task based on classical Navon figures [22]. Local-global stimuli were comprised of four shapes – a circle, X, triangle, or square – that were used to build the global and local aspects of the stimuli. On all trials, the local feature did not match the global feature, ensuring that subjects could not use information about one scale to infer information about another. Stimuli were presented on a black background in a block design with three block types (See Fig. 2). In the first block type, subjects viewed white local-global stimuli. In the second block type, subjects viewed green local-global stimuli. In the third block type, stimuli switched between white and green across trials uniformly at random with the constraint that 70% of trials included a switch in each block. In all blocks, subjects were instructed to report only the local features of the stimuli if the stimulus was white and to report only the global feature of the stimuli if the stimulus was green. Blocks were administered in a random order. Subjects responded using their right hand with a four-button box. All subjects were trained on the task outside the scanner until proficient at reporting responses using a fixed mapping between the shape and button presses (i.e., index finger = "circle", middle finger = "X", ring finger = "triangle", pinky finger = "square"). In the scanner, blocks were administered with 20 trials apiece separated by 20 s fixation periods with a white crosshair at the center of the screen. Each trial was presented for a fixed duration of 1900 ms separated by an interstimulus interval of 100 ms during which a black screen was presented. 13 Diffusion spectrum imaging acquisition and processing Diffusion spectrum images (DSI) were acquired on a Siemens 3.0T Tim Trio for all subjects along with a T1-weighted anatomical scan at each scanning session. We followed a parallel strategy for data acquisition and construction of streamline adjacency matrices as in previous work [25, 39]. DSI scans sampled 257 directions using a Q5 half-shell acquisition scheme with a maximum b-value of 5,000 and an isotropic voxel size of 2.4 mm. We utilized an axial acquisition with the following parameters: repetition time (TR) = 5 s, echo time (TE)= 138 ms, 52 slices, field of view (FoV) (231, 231, 125 mm). We acquired a three-dimensional SPGR T1 volume (TE = minimal full; flip angle = 15 degrees; FOV = 24 cm) for anatomical reconstruction. All subjects volunteered with informed consent in writing in accordance with the Institutional Review Board/Human Subjects Committee, University of Pennsylvania. DSI data were reconstructed in DSI Studio (www.dsi-studio.labsolver.org) using q-space diffeomorphic reconstruction (QSDR)[40]. QSDR first reconstructs diffusion-weighted im- ages in native space and computes the quantitative anisotropy (QA) in each voxel. These QA values are used to warp the brain to a template QA volume in Montreal Neurological Institute (MNI) space using the statistical parametric mapping (SPM) nonlinear registration algorithm. Once in MNI space, spin density functions were again reconstructed with a mean diffusion distance of 1.25 mm using three fiber orientations per voxel. Fiber tracking was performed in DSI studio with an angular cutoff of 35◦, step size of 1.0 mm, minimum length of 10 mm, spin density function smoothing of 0.0, maximum length of 400 mm and a QA threshold determined by DWI signal in the colony-stimulating factor. Deterministic fiber tracking using a modified FACT algorithm was performed until 1,000,000 streamlines were reconstructed for each individual. Anatomical scans were segmented using FreeSurfer[41] and parcellated using the connec- tome mapping toolkit [42]. A parcellation scheme including n = 129 regions was registered to the B0 volume from each subject's DSI data. The B0 to MNI voxel mapping produced via QSDR was used to map region labels from native space to MNI coordinates. To extend region labels through the grey-white matter interface, the atlas was dilated by 4 mm [43]. Dilation was accomplished by filling non-labelled voxels with the statistical mode of their neighbors' labels. In the event of a tie, one of the modes was arbitrarily selected. Each streamline was labeled according to its terminal region pair. Finally, we included a cerebellar parcellation [24]. We used FSL to nonlinearly register the individual's T1 to MNI space. Then, we used the inverse warp parameters to warp the 14 cerebellum atlas to the individual T1. We registered the subject's DSI image to the T1. We used the inverse parameters from this registration to map the individualized cerebellar parcels into the subject's DSI space. Finally, we merged the cerebellar label image with the dilated cortical and subcortical parcellation image. From these data and parcellation, we constructed an anatomical connectivity matrix, A whose element Aij represented the number of streamlines connecting different regions [17], divided by the sum of volumes for regions i and j [44]. Prior to data analysis, all cerebellum- to-cerebellum edges were removed from each individual's matrix because cerebellar lobules are demonstrably not anatomically connected directly to one another [45]. Functional imaging acquisition and processing fMRI images were acquired during the same scanning session as the DSI data on a 3.0T Siemens Tim Trio whole-body scanner with a whole-head elliptical coil by means of a single- shot gradient-echo T2* (TR = 1500 ms; TE = 30 ms; flip angle = 60◦; FOV = 19.2 cm, resolution 3mm x 3mm x 3mm). Preprocessing was performed using FEAT v. 6.0 (fMRI Expert Analysis Tool) a component of the FSL software package [46]. To prepare the func- tional images for analyses, we completed the following steps: skull-stripping with BET to remove non-brain material, motion correction with MCFLIRT (FMRIB's Linear Image Reg- istration Tool; [46]), slice timing correction (interleaved), spatial smoothing with a 6-mm 3D Gaussian kernel, and high pass temporal filtering to reduce low frequency artifacts. We also performed EPI unwarping with fieldmaps in order to improve subject registration to stan- dard space. Native image transformation to a standard template was completed using FSL's affine registration tool, FLIRT [46]. Subject-specific functional images were co-registered to their corresponding high-resolution anatomical images via a Boundary Based Registration technique (BBR [47]) and were then registered to the standard MNI-152 structural template via a 12-parameter linear transformation. Finally, each participant's individual anatomical image was segmented into grey matter, white matter, and CSF using the binary segmentation function of FAST v. 4.0 (FMRIB's Automated Segmentation Tool [48]). The white matter and CSF masks for each participant were then transformed to native functional space and the average timeseries were extracted. Images were spatially smoothed using a kernel with a full-width at half-maximum of 6 mm. These values were used as confound regressors on our time series along with 18 translation and rotation parameters as estimated by MCFLIRT [49]. 15 Functional decomposition into anatomical networks We analyze the signal defined on a connected, weighted, and symmetric graph, G = (V, A), where V = {1, . . . , n} is a set of n vertices or nodes representing individual brain regions and A ∈ Rn×n is defined as above. Because the network A is symmetric, it has a complete set of orthonormal eigenvectors associated with it [50, 51]. For this reason, it has an eigenvector decomposition, A = VΛVT , in which Λ is the set of eigenvalues, ordered so that λ0 ≤ λ1 ≤ . . . ≤ λn−1, and V = {vk}n−1 k=0 is the set of associated eigenvectors. Following [51, 52], we use the eigenvector matrix to define the Graph Fourier Transform (GFT) of the graph signal x ∈ Rn, defined as x = VT x. (1) Given x = [x0, . . . , xn−1]T , we can express our original signal as x =(cid:80)n−1 k=0 xkvk, a sum of the eigenvector components vk. The contribution of vk to the signal x is the GFT compo- nent xk. Note that the smoothness of vk on the network can be evaluated in the quadratic k Avk = (cid:80) form vT composition. The quantity (cid:80) i,j∈V Aijvk(i)vk(j) and that vT k Avk = λk is given by the eigenvector de- i,j∈V Aijvk(i)vk(j) will be negative when the signal is varied (highly connected regions possess signals of different signs), and positive when the signal is smooth (highly connected regions possess signals of same signs); for these reasons, this quan- tity can be thought of as a measure of smoothness (alignment). Consequently, these GFT coefficients xk for small values of k indicate how much variables that are highly misaligned (liberality) with structure contribute to the observed brain signal x. GFT coefficients xk for large values of k describe how much signals that are aligned with the anatomical network contribute to the observed brain signal x. The inverse (i)GFT of x with respect to A is defined as x = Vx. (2) Given a graph signal x with GFT x, we can isolate the liberal components corresponding to the lowest KL eigenvectors by applying a graph filter HL that only keeps components with k < KL and sets other components to 0. The signal xL then contains the "liberal" components of x (those with a low alignment with network structure). Apart from the graph low-pass filter HL, we also consider a middle graph regime HM, which keeps only components 16 in the range of KL ≤ k < n−KA, and an "aligned" graph regime HA, such that only network- aligned components with n− KA ≤ k are kept. Therefore, the liberal regime takes the lowest KL components, the alignment regime takes the highest KA components, and the middle regime captures the middle n − KL − KM components (here, we use the components with the 10 lowest alignment values to represent the liberal regime and the components with the 10 highest alignment values to represent the aligned regime; see Supplement for robustness of results to parameter selection). As such, if we use xM and xA to respectively denote the signals represented by the middle and highly aligned regimes, we have that the original signal can be written as the sum x = xL + xM + xA. This formulation gives a decomposition of the original signals x into liberal, moderately aligned, and highly aligned components that respectively represent signals that have high, medium, and low signal deviation with respect to the anatomical connectivity between brain regions. Prior work has consistently demonstrated that the aligned and liberal components aid in better estimation of unknown movie ratings in recommendation systems [53], better predic- tion of cancer using gene interaction networks [54, 55], and learning in neuroimaging data, where learning-related processes are demonstrably expressed in low and high components in fMRI data, and where the middle component xM is demonstrably less reliable and behav- iorally uninformative [56]. In the supplement, we perform a similar analysis to [56] but with the current data to examine the stability of our low and high alignment measurements to parameter selection. The data indicates that the low and high alignment components in the current data are stable. Mathematically, this is expected in general in applications of the current approach because eigenvalues at the extreme low and high end are isolated from the middle values, which leads to robustness in the high and low ranges of the decomposition [57]. We note that this approach allows signals on the anatomical network to contain both aligned and liberal components represented in the same region at a single TR. This flexibility occurs because the anatomical network of n nodes has n2 entries (i.e., connection information is encoded in the anatomical adjacency matrix for any node i to any node j). Rather than examining a single BOLD signal measurement as n independently observed values, the GFT considers the signal to be a composite of contributions to the signal across subject's anatomical network topology. The decomposition occurs across the entire set of signals (here, the vector of BOLD magnitudes across regions at a single TR), where there are only n entries. The GFT applied here leverages the fact that the n entries in a given vector are not 17 Figure 5: Signal decomposition into anatomy. BOLD signals are decomposed into aligned and liberal signal components. Left of equation: a schematic BOLD signal on a simple anatomical network. Here, two signals are stronger in the high direction than the low direction. Right of equation: the signals across the network are decomposed into an aligned and liberal component. The original signals can be reconstructed from a basis set including a weighted part of the signal that is aligned with the anatomical network and another part that is liberal with respect to the anatomical network. isolated, but are signals on top of the complex anatomical network. In the current approach, instead of focusing on the single BOLD value observed at each region as a discrete entity, the decomposition is sensitive to the observation of pairwise differences among BOLD signals relative to that expected by the anatomical network. Some portion of each given region's BOLD signal is estimated to be liberal with respect to the network, which is represented by xL, and some portion is estimated to be aligned with the network, which is represented by xA (See Fig. 5). This mathematical separation establishes the notions of alignment and liberality of the BOLD signals in the anatomical network. All individual regions in the brain will have some degree of alignment and some degree of liberality given the complexity of BOLD signal patterns across the network, unless the observed BOLD signals in all regions are perfectly aligned or perfectly misaligned with the subject's anatomical network. This highlights an important strength of the use of the Graph Fourier Transform to examine functional signal liberality in anatomical brain networks: the signal can be understood as a network level composite of aligned and liberal signals, and the extent to which individual regions contribute to these properties can be examined as the variation in the weights of the region's contribution to each of the aligned and liberal components. 18 =+xAxLxBOLD measurementAlignedLiberal Relating signals to behavior Following the signal decomposition into aligned and liberal signals, we associated the sig- nal concentrations with median switch cost (median response time during switching trials versus no-switching trials) performance for all accurate trials. To do so, we computed a partial Pearson's correlation between the observed signal value for each subject with their median switch cost using the average framewise displacement across BOLD measurements as a second-level control for the influences of motion. Specifically, to examine the relationship between alignment and switch costs across subjects, we computed the partial correlation for the mean of xA for each subject with subjects' switch costs, controlling for average frame- wise displacement. Then, to examine the relationship between liberality and switch costs, across subjects, we computed the partial correlation for the mean of xL for each subject with subjects' switch costs, controlling for average framewise displacement. We additionally repeated these analyses including age and sex and found similar slopes of the associations between the liberality values and switch costs (see Supplement). System permutation test To examine the spatial significance of system-level concentration of aligned and liberal signals, we performed a non-parametric permutation test for each signal class. Separately for each of xL and xA, we shuffled the observed mean signal concentration values across regions in 10,000 permutations for aligned and liberal signals and computed a null distribution of system mean signal concentrations for each system. Signals were judged to be significantly concentrated in a system if the mean signal concentration in the system was greater or less than 95% of the null permutations. 19 Supplementary Information and Analyses Behavior Across the 28 subjects, individuals were 94% (St.D. = 1%) accurate across all trials. The average median response time for accurate non-switching trials was 0.89 s (St.D. = 0.12 s). The average median response time for accurate switching trials across subjects was 1.22 s (St.D. = 0.18 s). The average switch cost across subjects was 0.32 s (St.D. = 0.08 s). Signal alignment as a trait-level variable in human brain networks In the main text, we report intraclass correlations that indicate that alignment and liberality form "trait" variables due to the consistency of aligned and liberal signal patterns regardless of the task period during which the signals were sampled. To examine the stability of aligned and liberal signals across all individual conditions, we compute a Pearson's correlation coefficient between the signal concentration vector across regions for each condition pair among aligned and liberal signal vectors for the fixation, no switch, and switching blocks. We observe that the aligned and liberal signals are very highly stable for each pairwise correlation within each signal type, but only moderately correlated across signal types (within aligned: mean R = 0.99, p (cid:28) 0.001, within liberal: mean R = 0.99, p (cid:28) 0.001, between aligned and liberal: mean R = 0.69, p (cid:28) 0.001; See. Fig. 1). These results establishe that during periods of resting fixation, no switching demands, and switching demands, alignment and liberality are highly stable subject-level traits. Signal variability in systems In the main text, we note that signals that are highly aligned with the underlying anatom- ical networks were concentrated in fronto-parietal, cingulo-opercular, default mode, and subcortical systems. This is especially interesting given numerous findings identifying the fronto-parietal and cingulo-opercular systems to contain variable signals over time and more flexible changes in network organization during both rest and task performance [32, 26]. To examine the BOLD signal variability in systems across TRs in the current study, we com- pute the average standard deviation of BOLD time series values over trials for all regions within each system. Then, we conduct a non-parametric permutation test in which we shuf- fle the assignment of regions to systems (n = 10, 000) and compute a null distribution of 20 Figure 6: Aligned and liberal signals are consistent traits in human brain networks. Correlations among conditions decomposing BOLD fMRI signals using the eigenspectrum of indi- viduals' anatomical networks. Aligned and liberal signals are highly stable at the subject level. Ali = Aligned, Lib = Liberal. average time series standard deviations over TRs for each system. We judge a system to be significantly variable or invariant over time if its observed average standard deviation is greater than or less than 97.5% of null permutations. We find that the fronto-parietal and cingulo-opercular systems are more variable across TRs than the null expectation, whereas the ventral attention, somatosensory, and cerebellar systems are less variable. Considered together with our main results examining aligned and liberal signal concentrations, these re- sults indicate that in cognitive control systems that are functionally dynamic, signal variance is strongly organized by aligning with subject's anatomical networks from TR to TR. System flexibility is associated with signal alignment In the main manuscript, we focus on the trial-to-trial signal alignment with anatomy across the brain as a basic spatiotemporal unity contributing to brain signal organization. We re- mark that previous analyses have focused on a notion of "flexibility" in dynamic networks. This notion is represented by creating a three-dimensional tensor from consecutive or over- lapping association matrices (e.g., using correlation, coherence, or mutual information) and observing the changes in node community allegiance over time. Flexible nodes are those that exhibit a relatively high degree of changes in community assignments. Distinct from the cur- rent analysis, flexibility occurs at a longer temporal scale since tensors are constructed from 21 Lib Fixation Lib No Switch Lib Switch Ali Switch Ali No Switch Ali Fixation Pearson's Correlation Coefficient 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 Figure 7: System-level variability across TRs. Plot of system-level variability in BOLD time series across TRs. Boxes and whiskers illustrate the distribution of system average standard devi- ation of BOLD time series over TRs for each system. Green circles illustrate the observed average standard deviations across TRs in each system. Asterisks denote systems in which the observed means were higher or lower than 97.5% of permuted values. many adjacency matrices over time, which in turn are computed over many TRs of time series data. In this context, it is interesting to consider whether the anatomical alignment of BOLD signals at the TR level is associated with the expression of flexibility at the temporal network level. To explore the relationship between flexible systems and signal alignment with anatomy, we explicitly computed flexibility and associated these values within each participant's align- ment values. Specifically, we followed [58] to examine the dynamic community structure of the current data. We constructed a multilayer temporal network for each subject by com- puting the correlation between all region pairs. Correlations whose significance did not pass false discovery rate correction with a threshold of 0.05 were set to zero [59] for each of 10 non- overlapping windows (40 TRs apiece) over the entire task session. We maximized multilayer modularity Qml following [60]: (cid:88) ijlr (cid:26)(cid:18) Aijl − γl kilkjl 2ml (cid:19) (cid:27) δlr + δijCjlr δ(gil, gjr), Qml = 1 2µ where the adjacency matrix of layer l (i.e., time window number l) has components Aill, γl is the resolution parameter of layer l, gil gives the community assignment of node i in layer 22 - subcortical cingulo-opercular default mode dorsal attention fronto-parietal somatosensory auditory ventral attention visual cerebellum other * * * * * System signal standard deviation over TRs l, gjr gives the community assignment of node j in layer r, Cjlr is the connection strength between node j in layer r and node j in layer l (see the discussion below), kil is the strength of node i in layer l, 2µ =(cid:80) jr kjr, kjl = kjl + cjl, and cjl =(cid:80) r Cjlr. For simplicity, we set the resolution parameter γl to unity and we have set all non-zero Cjlr to a constant C, which we term the inter-layer coupling. Here, we compute community assignments for C = 1. For each subject, we optimized the quality function 100 times and identified a representative consensus partition for all nodes over all windows [61]. Within each subject's consensus partition, we defined the flexibility of a node fi to be the number of times that node changed modular assignment throughout the session, normalized by the total number of changes that were possible (i.e., by the number of consecutive pairs of layers in the multilayer framework) [59]. We then defined the flexibility of each system examined in the main text as the mean flexibility over all nodes in the subsystem. Finally, we calculated the Pearson correlation coefficient between the mean system flexibility and mean system alignment. Notably, we found that mean flexibility and mean alignment were significantly positively correlated with one another across systems (R = 0.15, p = 0.015). This result indicates that trial-level signal alignment across the brain is positively associated with the expression of temporal system flexibility over much larger temporal scales across brain networks. Signal alignment in anatomy is similar to function To provide further intuition for our observation that signal alignment is greatest in the most flexible systems, we highlight that the seemingly counterintuitive observation – the regions whose signals are aligned with the underlying structural network are also the regions that display the greatest network flexibility across time points – is mathematically justifiable. The reason comes from the close relationship between the graph Fourier transform (GFT) and Principal Component Analysis (PCA) [62, 63], when the network represents pairwise correlations between brain regions across time points. In this case, from PCA, the signal distribution that explains most of the variation of signals across time points is represented by the eigenvectors associated with large eigenvalues of the correlation matrix (functional connectivity network). Simultaneously, recall from GFT that the eigenvectors associated with large eigenvalues of the correlation matrix depict signal distributions that align well with the network [51, 53]. This implies that the signal distribution that aligns with the network representing pairwise correlations between brain regions also varies the most across 23 time points. Stated simply, it can be derived mathematically that the most flexible systems are most aligned with the functional network constructed from those signals [56]. Thus, crucial to the current result, if the correlation matrix's eigenspectrum is similar to the anatomical network, the most anatomically aligned functional signals will also be related to the most functionally aligned and flexible signals. In order to evaluate this relationship empirically, we examined the correlation between signals aligned with the structural networks and signals aligned with the functional networks. For each subject in the experiment, we constructed the subject's correlation matrix across all TRs and performed the same decom- position of each BOLD TR from our primary analysis using each subject's (1) anatomical network and then repeated this analysis using (2) the correlation matrix across all BOLD TRs. Next, we examined the similarity between the signals aligned with the functional cor- relation network and those aligned with the anatomical network. A significant correlation would indicate that the decomposition of the aligned portion of signals is organized with the functional correlation network similarly to its decomposition into the anatomical network, suggesting that the relationships between flexibility and alignment is expressed both in the subject's anatomical and mean functional correlation network. We find a strong correlation between the anatomically and functionally aligned signals (R = 0.814, p = 1.613 × 10−27). This empirical relationship indicates that signals most aligned with the functional network are those most aligned with the anatomical network. Thus, the eigenspectrum of the functional network that contributes to the tendency for flexi- ble functional signals to align is related to the anatomical eigenspectrum cross subjects. This highlights a fundamental relationship guiding BOLD signal dynamics in anatomical and func- tional network analysis. Both anatomical and functional correlation networks demonstrate an eigenspectrum that organizes flexible activity in the human brain. 24 Aligned and liberal signal associations with behavior across task conditions In the main text, we focus on the relationship between variability in signal alignment and variability in switch costs due to our emphasis on cognitive control in human brain networks. For reference, here we include a full table of results associating median response times in each condition (no switching, switching, and switch costs) with aligned and liberal signals across the 28 subjects. The results indicate that within the current signal decomposition using anatomical networks, only the liberal signals demonstrate a relationship with behavioral variability and specifically with switch costs. Table 1: Pearson's partial correlation coefficients: all median response time associa- tions with signal liberality and alignment. Behavior Lib Fix Lib NS Lib S Ali Fix Ali NS Ali S No Switch 0.2223 0.3503 0.2207 0.0328 0.1134 -0.0907 Switch 0.4264 0.5563 0.4026 0.1962 0.2298 0.0090 Switch Cost 0.6181 0.7137 0.5669 0.3912 0.3431 0.1599 Pearson's correlation coefficients using average framewise displacement as a covariate. Rows contain partial correlation coefficients associating each median response time at the subject level with their graph alignment. Columns represent the signal type including the fixation, no switch, and switch block signals. Lib = Liberal, Ali = Aligned, Fix = fixation, NS = no switch, S = switch. Table 2: Pearson's partial correlation p-values: all median response time associations with signal liberality and alignment. Behavior Lib Fix Lib NS Lib S Ali Fix Ali NS Ali S No Switch 0.2650 0.0733 0.2686 0.8709 0.5734 0.6526 Switch 0.0266 0.0026 0.0374 0.3267 0.2488 0.9644 Switch Cost 0.0006 <0.0001 0.0020 0.0436 0.0798 0.4258 Rows contain p-values corresponding to correlation coefficients in Table 1. Columns represent the signal type including the fixation, no switch, and switch block signals. Lib = Liberal, Ali = Aligned, Fix = fixation, NS = no switch, S = switch. 25 Inclusion of age and sex as covariates In the main text, we relate alignment and liberality to behavior using subjects' average framewise displacement as covariates. To examine whether our results are stable when accounting for subjects' age and sex, we recompute partial correlations for each task condition and alignment and liberality values using average framewise displacement, age, and sex as covariates. Our results remain significant with similar correlation values across conditions when including these covariates. Table 3: Pearson's partial correlation coefficients: all median response time associa- tions with signal liberality and alignment. Behavior Lib Fix Lib NS Lib S Ali Fix Ali NS Ali S No Switch 0.3508 0.4990 0.4116 0.1168 0.1842 -0.0042 Switch 0.5009 0.6443 0.5390 0.2319 0.2390 0.0527 Switch Cost 0.6028 0.7052 0.5984 0.3459 0.2629 0.1237 Pearson's correlation coefficients using average framewise displacement, age, and sex as co- variates. Rows contain partial correlation coefficients associating each median response time at the subject level with their graph liberality. Columns represent the signal type includ- ing the fixation, no switch, and switch block signals. Lib = Liberal, Ali = Aligned, Fix = fixation, NS = no switch, S = switch. Table 4: Pearson's partial correlation p-values: all median response time associations with signal liberality and alignment. Behavior Lib Fix Lib NS Lib S Ali Fix Ali NS Ali S No Switch 0.0856 0.0111 0.0409 0.5783 0.3780 0.9842 Switch 0.0107 0.0005 0.0054 0.2646 0.2499 0.8023 Switch Cost 0.0014 0.0001 0.0016 0.0903 0.2042 0.5559 Rows contain p-values corresponding to correlation coefficients in Table 3. Columns represent the signal type including the fixation, no switch, and switch block signals. Lib = Liberal, Ali = Aligned, Fix = fixation, NS = no switch, S = switch. 26 Robustness of aligned and liberal signals to parameter selection In the main text, we report results where we represent the liberal signals xL and aligned signals xH using cut points for KL and KH of the ten lowest (of KL) and highest (KH) values. To test the robustness of aligned and liberal signals to variations in these parameters, we decompose the observed BOLD signals into xL and xH under the choice of KL and KH from five below to five above the choice of parameters studied in the main manuscript. We then compute the correlation coefficient among xL and xH across all TRs for all subjects to examine the stability of aligned and liberal signals. We find that the observed signal decomposition is stable across parameter choices: for xL (liberal signals), the values are correlated at a mean of R = 0.95 (St.D. 0.03); for xH (aligned signals), the values are correlated at a mean of R = 0.99 (St.D. 0.02). Thus, the signals examined in the main text are highly stable over the choice of definition for the range of aligned and liberal signals, and especially for liberal signals. Null permutation test for the significance of the anatomical orga- nization of signals In the current analysis, the aligned signals are found to be the most concentrated in systems including those known to be functionally flexible in prior studies [32, 26]. It is interesting to examine the complexity of anatomical contributions to these signals. We consider whether anatomically aligned signals are organized by the simple contributions of nearest neighbor nodes in anatomical networks (i.e., only the nodes that share direct connections to each node) versus the specific configuration of the entire anatomical network. This allows us to identify whether the BOLD signals inclusive of the most flexible systems in the brain are aligned with anatomy as a function of the contribution of the entire network, which would potentially suggest that functionally flexible systems are organized by weighted contributions from the entire brain. To test this possibility, we conduct a non-parametric permutation test. Specifically, we randomize each individual's network 2000 times preserving the strength and degree sequence of nodes in the network. Then, we decompose the observed BOLD value for each region into its corresponding node in the randomized network across all TRs. We then calculate over subjects the average variance in original BOLD signals accounted for by the eigenvectors associated with the aligned signals for each random permutation. We summarize these 27 values as the mean variance accounted for across all null permutations and compare it to the mean variance accounted for by the original anatomical networks. We observe that the variance accounted for in the real networks exceeds that accounted for in 100% of the null permutations. This indicates that the BOLD signals are organized by anatomical features at a greater topological length across the networks than the direct connections among nodes (See Fig. 9). This finding indicates that signals that depend on underlying anatomy represent complex contributions from anatomical network topology. The behavioral relevance of this anatomically-aligned organization may be established in future studies. In particular, when the underlying network is constructed from fMRI measurements (functional similarity networks constructed from correlations), it is known that the components of BOLD signals that are the most aligned to the network also explain the most variation across time [56]. Here also we find that signals aligned with anatomical connectivity contain BOLD measurements that are the most variable over time in cingulo- opercular and fronto-parietal systems. Functional connectivity (e.g., correlation, coherence, and mutual information) networks evince similar but not completely overlapping topology when compared to anatomical networks [64]. Thus, analyzing the alignment of measured cor- relations on anatomical networks may provide an important extension to understand signals in the time domain in the human connectome. Mathematically, the same phenomenon for functional as well as anatomical networks implies that the eigenspectrum of the anatomical network is not that different from functional connectivity. Speculatively, this could provide a meaningful and efficient method to reliably recover anatomical networks from functional connectivity networks. Alternate measures of function and anatomy Do anatomy-decomposed functional BOLD signals better related to switch costs than other measures with similar intuitions? To test this possibility, we consider four comparison anal- yses involving the subcortical system from which the behaviorally relevant signals in our primary analyses originated. First, we compute the mean BOLD signal variance over all sub- cortical regions for each TR. This correlation represents a measure of differences in BOLD signals at each time measurement without considering its anatomical network context. We then relate the average BOLD signal variance over TRs to switch costs and find no signifi- cant relationship (R = −0.09, p = 0.68). Second, we compute the mean node strength – the sum of connections to the node – for each region in the subcortical system. This represents 28 Figure 8: Aligned signals are organized by topologically distant anatomical organization. Left: in the null permutations where the edges of the anatomical networks were randomized, a mean of 0.455% variance is accounted for by the eigenvectors corresponding to the aligned parts of the signal. In the observed data, a mean of 0.547% of variance is accounted for by the anatomical networks in their true configurations. the overall influence of the subcortical regions in the broader anatomical network. We cal- culate the Pearson correlation coefficient between the mean subcortical node strength and switch costs across subjects, and we find no significant relationship (R = −0.15, p = 0.48). Finally, within each subject, we compute the mean Pearson correlation coefficient between each BOLD signal TR and node strengths within the subcortical system, then we compute the Pearson correlation coefficient between these values and switch costs across subjects, finding no significant relationship (R = 0.17, p = 0.41). Taken together, these results demonstrate that simple measures of BOLD variation, local anatomical network influences, and the relationship between the two are insufficient to account for switch cost variability over subjects. Indeed, decomposing signals in the context of the entire anatomical network provides superior associative value for the cognitive control measure. 29 Number of null observations % variance explained Null Real 0 20 40 60 80 100 120 140 160 180 0.42 0.44 0.46 0.48 0.50 0.52 0.54 0.56 Figure 9: Alternate measures of framewise BOLD variance, anatomy, and function- anatomy relationships in subcortical systems do not associate with switch costs. (A) The relationship between mean BOLD variance in the subcortical system and switch costs. (B) The relationship between mean anatomical node degree in the subcortical system and switch costs. (C) The relationship between the correlation between BOLD signals and anatomical node degree across regions in the subcortical systems and switch costs. Anatomical network null model for behavioral correlations with switch costs. To examine the importance of the specific anatomical configuration of brain networks for the association between liberal signals and switch costs, we performed a null permutation test. Specifically, we performed 200 permutations in which we generated a random network preserving the strength and degree distributions of the original anatomical networks for each subject. Then, we repeated the BOLD signal decomposition into the null anatomical network for each subject and correlated the liberal signals with switch costs in each permutation to generate a null distribution of correlations. From this null distribution, we computed the proportion of null correlation values greater than the observed correlation when using the real anatomical networks. We find that the observed correlation is greater than 99% of correlations (Fig. 10). This result indicates that the specific configuration of the true anatomical network drives the correlation between liberal signal alignment and switch costs. 30 01000200030000.20.250.30.350.40.450.5500100015000.20.250.30.350.40.450.5-0.04-0.0200.020.040.20.250.30.350.40.450.5Switch cost (s)Switch cost (s)Switch cost (s)Mean BOLD varianceMean node anatomical strengthMean R between BOLD, anatomical strengthABC Figure 10: A null permutation test demonstrates that specific anatomical network or- ganization is required to identify behaviorally relevant BOLD signals. Blue bars in the histogram illustrate the Pearson's correlation coefficients between switch costs and liberal signals over null permutations. Orange line designates the observed correlation value of the real network. Relationships between motion and signal alignment. In the results reported in the main text and supplement, we included a motion covariate (average framewise displacement across BOLD TRs) in all analyses. Motion is not signifi- cantly correlated with switch cost (R = 0.26, p = 0.16), but is correlated significantly with liberal signals (R = 0.77, p = 1.10×10−6), justifying its inclusion as a covariate in behavioral analyses. Importantly, when motion is not included as a covariate in the partial correlation between liberal signals and switch costs, the correlation (R = 0.59 , p = 0.001) is highly similar to that observed when including motion as a covariate (R = 0.57, p = 0.002). This finding indicates that the behaviorally relevant portion of the liberal signals is not driven by motion. References [1] Rogers, R. D. & Monsell, S. Costs of a predictible switch between simple cognitive tasks. Journal of experimental psychology: General 124, 207 (1995). [2] Szczepanski, S. M. & Knight, R. T. Insights into human behavior from lesions to the prefrontal cortex. Neuron 83, 1002–1018 (2014). 31 Number of null observations Pearson's R 0.20.30.40.50.60.7051015202530Null Real [3] Clark, L. R. et al. Specific measures of executive function predict cognitive decline in older adults. Journal of the International Neuropsychological Society 18, 118–127 (2012). [4] Richland, L. E. & Burchinal, M. R. Early executive function predicts reasoning devel- opment. Psychological science 24, 87–92 (2013). [5] Davis, J. C., Marra, C. A., Najafzadeh, M. & Liu-Ambrose, T. The independent con- tribution of executive functions to health related quality of life in older women. BMC geriatrics 10, 16 (2010). [6] Gunaydin, L. A. & Kreitzer, A. C. Cortico-basal ganglia circuit function in psychiatric disease. Annual review of physiology 78, 327–350 (2016). [7] Casey, B. et al. Early development of subcortical regions involved in non-cued attention switching. Developmental science 7, 534–542 (2004). [8] Cole, M. W. et al. Multi-task connectivity reveals flexible hubs for adaptive task control. Nature neuroscience 16, 1348–1355 (2013). [9] Heyder, K., Suchan, B. & Daum, I. Cortico-subcortical contributions to executive control. Acta psychologica 115, 271–289 (2004). [10] Luk, G., Green, D. W., Abutalebi, J. & Grady, C. Cognitive control for language switching in bilinguals: A quantitative meta-analysis of functional neuroimaging studies. Language and cognitive processes 27, 1479–1488 (2012). [11] Hikosaka, O. & Isoda, M. Switching from automatic to controlled behavior: cortico- basal ganglia mechanisms. Trends in cognitive sciences 14, 154–161 (2010). [12] Hosoda, C., Hanakawa, T., Nariai, T., Ohno, K. & Honda, M. Neural mechanisms of language switch. Journal of Neurolinguistics 25, 44–61 (2012). [13] Leunissen, I. et al. Subcortical volume analysis in traumatic brain injury: the impor- tance of the fronto-striato-thalamic circuit in task switching. Cortex 51, 67–81 (2014). [14] Yehene, E., Meiran, N. & Soroker, N. Basal ganglia play a unique role in task switching within the frontal-subcortical circuits: evidence from patients with focal lesions. Journal of Cognitive Neuroscience 20, 1079–1093 (2008). 32 [15] Sporns, O., Tononi, G. & Kotter, R. The human connectome: A structural description of the human brain. PLoS Computational Biology 1, e42 (2005). [16] Alstott, J., Breakspear, M., Hagmann, P., Cammoun, L. & Sporns, O. Modeling the impact of lesions in the human brain. PLoS computational biology 5, e1000408 (2009). [17] Hermundstad, A. M. et al. Structural foundations of resting-state and task-based func- tional connectivity in the human brain. Proceedings of the National Academy of Sciences 110, 6169–6174 (2013). [18] Honey, C. J., Kotter, R., Breakspear, M. & Sporns, O. Network structure of cere- bral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences 104, 10240–10245 (2007). [19] Medaglia, J. D., Lynall, M.-E. & Bassett, D. S. Cognitive network neuroscience. Journal of cognitive neuroscience (2015). [20] Sporns, O. Contributions and challenges for network models in cognitive neuroscience. Nature Neuroscience 17, 652–660 (2014). [21] Power, J. D. et al. Functional network organization of the human brain. Neuron 72, 665–678 (2011). [22] Navon, D. Forest before trees: The precedence of global features in visual perception. Cognitive psychology 9, 353–383 (1977). [23] Cammoun, L. et al. Mapping the human connectome at multiple scales with diffusion spectrum mri. Journal of Neuroscience Methods 203, 386–397 (2012). [24] Diedrichsen, J., Balsters, J. H., Flavell, J., Cussans, E. & Ramnani, N. A probabilistic mr atlas of the human cerebellum. Neuroimage 46, 39–46 (2009). [25] Gu, S. et al. Controllability of structural brain networks. Nature Communications 6, 8414 (2015). [26] Braun, U. et al. Dynamic reconfiguration of frontal brain networks during executive cognition in humans. Proceedings of the National Academy of Sciences 112, 11678– 11683 (2015). 33 [27] Mayhew, S. D. et al. Global signal modulation of single-trial fmri response variability: Effect on positive vs negative bold response relationship. NeuroImage 133, 62–74 (2016). [28] Sekutowicz, M. et al. Striatal activation as a neural link between cognitive and percep- tual flexibility. NeuroImage 141, 393–398 (2016). [29] Liston, C., Matalon, S., Hare, T. A., Davidson, M. C. & Casey, B. Anterior cingulate and posterior parietal cortices are sensitive to dissociable forms of conflict in a task- switching paradigm. Neuron 50, 643–653 (2006). [30] Cole, M. W., Bassett, D. S., Power, J. D., Braver, T. S. & Petersen, S. E. Intrinsic and task-evoked network architectures of the human brain. Neuron 83, 238–251 (2014). [31] Watanabe, T., Masuda, N., Megumi, F., Kanai, R. & Rees, G. Energy landscape and dynamics of brain activity during human bistable perception. Nature communications 5 (2014). [32] Mattar, M. G., Betzel, R. F. & Bassett, D. S. The flexible brain. Brain 139, 2110–2112 (2016). [33] Miyake, A., Friedman, N. P., Emerson, M. J., Witzki, A. H. & Howerter, A. The unity and diversity of executive functions and their contributions to complex "frontal lobe" tasks: A latent variable analysis. Cognitive Psychology 41, 49–100 (2000). [34] Fedorenko, E. The role of domain-general cognitive control in language comprehension. Frontiers in psychology 5, 335 (2014). [35] Hermundstad, A. M. et al. Structurally-constrained relationships between cognitive states in the human brain. PLoS Comput Biol 10, e1003591 (2014). [36] Belleville, S., Bherer, L., Lepage, ´E., Chertkow, H. & Gauthier, S. Task switching capacities in persons with alzheimer's disease and mild cognitive impairment. Neu- ropsychologia 46, 2225–2233 (2008). [37] Kehagia, A. A., Barker, R. A. & Robbins, T. W. Neuropsychological and clinical heterogeneity of cognitive impairment and dementia in patients with parkinson's disease. The Lancet Neurology 9, 1200–1213 (2010). [38] Kinnunen, K. M. et al. White matter damage and cognitive impairment after traumatic brain injury. Brain awq347 (2010). 34 [39] Betzel, R. F., Gu, S., Medaglia, J. D., Pasqualetti, F. & Bassett, D. S. Optimally controlling the human connectome: the role of network topology. Scientific Reports (2016). [40] Yeh, F.-C., Wedeen, V. J. & Tseng, W.-Y. I. Estimation of fiber orientation and spin density distribution by diffusion deconvolution. Neuroimage 55, 1054–1062 (2011). [41] Fischl, B. Freesurfer. Neuroimage 62, 774–781 (2012). [42] Cammoun, L. et al. Mapping the human connectome at multiple scales with diffusion spectrum mri. Journal of neuroscience methods 203, 386–397 (2012). [43] Cieslak, M. & Grafton, S. Local termination pattern analysis: a tool for comparing white matter morphology. Brain imaging and behavior 8, 292–299 (2014). [44] Hagmann, P. et al. Mapping the structural core of human cerebral cortex. PLoS Biol 6, e159 (2008). [45] Voogd, J. & Glickstein, M. The anatomy of the cerebellum. Trends in cognitive sciences 2, 307–313 (1998). [46] Jenkinson, M., Beckmann, C. F., Behrens, T. E., Woolrich, M. W. & Smith, S. M. Fsl. Neuroimage 62, 782–790 (2012). [47] Greve, D. N. & Fischl, B. Accurate and robust brain image alignment using boundary- based registration. Neuroimage 48, 63–72 (2009). [48] Zhang, Y., Brady, M. & Smith, S. Segmentation of brain mr images through a hid- den markov random field model and the expectation-maximization algorithm. IEEE transactions on medical imaging 20, 45–57 (2001). [49] Jenkinson, M., Bannister, P., Brady, M. & Smith, S. Improved optimization for the ro- bust and accurate linear registration and motion correction of brain images. Neuroimage 17, 825–841 (2002). [50] Chung, F. Spectral graph theory, vol. 92 (American Mathematical Soc., 1997). [51] Sandryhaila, A. & Moura, J. M. Discrete signal processing on graphs. IEEE transactions on signal processing 61, 1644–1656 (2013). 35 [52] Shuman, D. I., Narang, S. K., Frossard, P., Ortega, A. & Vandergheynst, P. The emerging field of signal processing on graphs: Extending high-dimensional data analysis to networks and other irregular domains. IEEE Signal Processing Magazine 30, 83–98 (2013). [53] Ma, J., Huang, W., Segarra, S. & Ribeiro, A. Diffusion filtering for graph signals and its use in recommendation systems. In Acoustics, Speech and Signal Processing (ICASSP), 2016 IEEE Int. Conf. on, 4563–4567 (Shanghai, China, 2016). [54] Segarra, S., Huang, W. & Ribeiro, A. Diffusion and superposition distances for signals supported on networks. Signal Inform. Process. over Network., IEEE Trans. on 1, 20–32 (2015). [55] Huang, W., Segarra, S. & Ribeiro, A. Diffusion distance for signals supported on networks. In Proc. Asilomar Conf. on Signals Systems Computers, 1219–1223 (Asilomar CA, 2015). [56] Huang, W. et al. Graph frequency analysis of brain signals. J. Sel. Topics Signal Process. 10, 1189–1203 (2016). [57] Spielman, D. Spectral graph theory. Lecture Notes, Yale University 740–0776 (2009). [58] Bassett, D. S. et al. Robust detection of dynamic community structure in networks. Chaos: An Interdisciplinary Journal of Nonlinear Science 23, 013142 (2013). [59] Bassett, D. S. et al. Dynamic reconfiguration of human brain networks during learning. Proceedings of the National Academy of Sciences 108, 7641–7646 (2011). [60] Mucha, P. J., Richardson, T., Macon, K., Porter, M. A. & Onnela, J.-P. Community structure in time-dependent, multiscale, and multiplex networks. Science 328, 876–878 (2010). [61] Bassett, D. S. & Siebenhuhner, F. Multiscale network organization in the human brain. Multiscale Analysis and Nonlinear Dynamics: From Genes to the Brain 179–204 (2013). [62] Leonardi, N. et al. Principal components of functional connectivity: a new approach to study dynamic brain connectivity during rest. NeuroImage 83, 937–950 (2013). [63] Viviani, R., Gron, G. & Spitzer, M. Functional principal component analysis of fmri data. Human brain mapping 24, 109–129 (2005). 36 [64] Bullmore, E. & Sporns, O. Complex brain networks: graph theoretical analysis of structural and functional systems. Nature Reviews Neuroscience 10, 186–198 (2009). 37 Acknowledgements JDM acknowledges support from the Office of the Director at the National Institutes of Health through grant number 1-DP5-OD-021352-01. DSB acknowledges support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Research Laboratory and the Army Research Office through contract numbers W911NF- 10-2-0022 and W911NF-14-1-0679, the National Institute of Health (2-R01-DC-009209-11, 1R01HD086888-01, R01-MH107235, R01-MH107703, R01MH109520, 1R01NS099348 and R21-M MH-106799), the Office of Naval Research, and the National Science Foundation (BCS-1441502, CAREER PHY-1554488, BCS-1631550, and CNS-1626008). The content is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies. Competing Interests The authors declare that they have no competing financial interests. Correspondence Correspondence and requests for materials should be addressed to D.S.B. (email: [email protected]). 38
1103.0668
1
1103
2011-03-03T12:37:57
Spike Onset Dynamics and Response Speed in Neuronal Populations
[ "q-bio.NC" ]
Recent studies of cortical neurons driven by fluctuating currents revealed cutoff frequencies for action potential encoding of several hundred Hz. Theoretical studies of biophysical neuron models have predicted a much lower cutoff frequency of the order of average firing rate or the inverse membrane time constant. The biophysical origin of the observed high cutoff frequencies is thus not well understood. Here we introduce a neuron model with dynamical action potential generation, in which the linear response can be analytically calculated for uncorrelated synaptic noise. We find that the cutoff frequencies increase to very large values when the time scale of action potential initiation becomes short.
q-bio.NC
q-bio
Spike Onset Dynamics and Response Speed in Neuronal Populations MPI for Dynamics and Self-Organization, Faculty of physics, Georg-August University, and Bernstein center for computational neuroscience, Gottingen, Germany Wei Wei and Fred Wolf Recent studies of cortical neurons driven by fluctuating currents revealed cutoff frequencies for action potential encoding of several hundred Hz. Theoretical studies of biophysical neuron models have predicted a much lower cutoff frequency of the order of average firing rate or the inverse membrane time constant. The biophysical origin of the observed high cutoff frequencies is thus not well understood. Here we introduce a neuron model with dynamical action potential generation, in which the linear response can be analytically calculated for uncorrelated synaptic noise. We find that the cutoff frequencies increase to very large values when the time scale of action potential initiation becomes short. PACS numbers: 87.19.ll, 05.40.-a, 87.19.ls Keywords: Action potential, Onset rapidness, Linear response, Cut-off frequency In the cerebral cortex of the brain information is en- coded in the action potential (AP) firing rates of a large ensemble of nerve cells. Recent experiments have ob- served a surprisingly high cutoff frequency for the action potential encoding of cortical neurons driven by fluctuat- ing input currents [1 -- 4]. In a seminal paper Kondgen et al. showed that the transmission function of layer 5 pyra- midal neurons for a noisy sinusoidal signal does not de- cay until about 200 Hz [1]. Later experiments confirmed such high cutoff frequencies for signals coded by both the mean current and noise strength [2] and in other types of cortical neurons [3, 4]. For an early observation of fast response see [5]. Previous theoretical studies of biophysi- cal neuron models, however, predicted cutoff frequencies of the order of the average firing rate or the inverse mem- brane time constant (below 20 Hz), much lower than the experimentally observed values [6 -- 8]. Thus, the origin of the high cutoff frequencies found in cortical neurons is currently not well understood. Numerical investigation of neuron models with dynamical AP generation, like the exponential integrate-and-fire (EIF) model or the gener- alized theta neurons, suggested that details of AP gen- eration can influence the dynamical response of neuronal populations [6 -- 9]. What is missing, however, is a trans- parent understanding of how and when the population cutoff frequency can dissociate from the basic single neu- ron timescale set by the mean firing rate and the time constant of membrane potential relaxation. In this work we present an analytically solvable model which explicitly describes the dynamical AP initiation process. A neuron initiates an AP if the membrane po- tential passes an unstable fixed point, the voltage thresh- old. In the leaky integrate-and-fire (LIF) model, for which the linear response is known analytically, the un- stable fixed point coincides with the absorbing bound- ary and a spike is triggered immediately when the mem- brane potential reaches this threshold [10, 11]. As a con- sequence, boundary induced artifacts dominate the re- sponse for high signal frequencies in the LIF model [6 -- 8]. One important advantage of our new model is that such boundary induced artifacts can be separated out mathe- matically, isolating the physically meaningful part of the response function. We first present the linear response for both encoding paradigms with white noise. We find that for a wide range of parameter settings the cutoff frequency is directly proportional to the AP onset rapid- ness for a noise coded signal. It therefore dissociates from the membrane time constant and can become arbitrarily large. For the mean current coded signal, however, the cutoff frequency is confined by the membrane time con- stant in the white noise case. We show by numerical simulation that this confinement can be broken when a finite correlation time in the synaptic noise is taken into account and high cutoff frequencies can be obtained for a large AP onset rapidness. Interestingly, experiments showed that the AP onset rapidness of cortical neurons is very large both in vitro and in vivo [12, 13], which may thus explain the occurrence of high cutoff frequen- cies. Our results provide a relationship between the spike onset dynamics and the population cutoff frequency that can be directly tested in physiological experiments. The simplest voltage dynamics that exhibits both a stable fixed point (the resting potential) and an unsta- ble fixed point (the voltage threshold) has a piecewise linear membrane current, composed of a leak current for low potential and a linear spike generating current for high potential [Fig. 1(a)]. The model is defined by the following Langevin equation τm v = −v + Θ(v − v0) (r + 1)(v − v0) + µ + ση(t) , (1) where v is the membrane potential relative to the rest- ing potential, τm is the membrane time constant, Θ(v) is the Heaviside step function, r is the AP onset rapid- ness which sets an effective time constant τm/r for the AP initiation process. The larger is r, the faster is the spike onset [Fig. 1(b)]. In biophysical models, the on- set rapidness will be set largely by intrinsic properties of the voltage dependent sodium channels, e.g., the gating ′ ′ )i = τmδ(t − t charge and the slope of the activation curve [12]. µ is the mean input current and σ is the amplitude of synaptic noise. η(t) is a Gaussian white noise satisfying hη(t)i = 0 ). The crossing point v0 of and hη(t)η(t the two pieces sets the rheobase current, which we use as the unit of voltage, v0 = 1. The threshold potential is vt = (1 + 1/r)v0. When the membrane potential reaches vb, the truncation point of the AP upstroke, it is reset to a voltage vr and stays there for an absolute refractory period τr. For convenience we take τm as the unit of time in analytical calculation. The Fokker-Plank equation corresponding to Eq. (1) has the following form ∂tP1 + ∂v(cid:18) − v + µ − ∂tP2 + ∂v(cid:18)r(v − vt) + µ − 1 2 σ2∂v(cid:19)P1 = 0 , 1 2 σ2∂v(cid:19)P2 = 0, (2) where P1(v, t) and P2(v, t) are the probability densities of membrane potential v for −∞ < v ≤ 1 and 1 < v ≤ vb respectively. The stationary firing rate can be found eas- ily when the boundary conditions are specified. From the reset assumption, we impose an absorbing boundary at vb and, as a result the probability density is zero there, P2(vb, t) = 0. The firing rate is given by the probabil- ity current at vb, ν(t) = − 1 2 σ2∂vP2(vb, t). At the reset voltage vr, the probability density is continuous, while its first derivative has a discontinuity from the reset con- dition: ∂vP1(v+ r , t) − ∂vP1(v−r , t) = ∂vP2(vb, t − τr). In addition, the density and its first derivative should be continuous at v = 1. When µ and σ are constants, the system is homogeneous and the stationary solution of Eq. (2), denoted as P01(v) and P02(v) respectively, can be found. The stationary firing rate ν0 is then obtained from the normalization condition of the density [14]. The stationary firing rate ν0 of the model Eq. (1)reduces to that for the LIF model for r → ∞. Figure 1(c) and D show the dependence of ν0 on the mean input µ and the amplitude of noise σ, respectively. The firing rate ν0 increases monotonically with r, µ, and σ and is relatively insensitive to r when r > 10. For the dynam- ical response, however, the r dependence is much more pronounced. The instantaneous firing rate of an ensem- ble of model neurons responds much faster for larger r to a step change in the noise level [14]. Linear response. -- When the input current to a neu- ron is weakly modulated, linear response theory can be applied to study the dynamical response properties of an ensemble of neurons. To this end, we consider a sinu- soidal signal ε cos(ωt), where ε is small. When the signal is encoded in the mean current, µ → µ + ε cos(ωt) or in the noise amplitude, σ → σ + εcos(ωt), the instantaneous firing rate can be written as ν(t) = ν0 + εν1c(ω)cos(ωt− φc(ω)) or ν(t) = ν0 + εσν1n(ω)cos(ωt − φn(ω)). Here ν1(ω) are the complex response functions. The absolute (a) v 1 τm r τm 0 1 vb v (b) ) t ( V 4 3 2 1 0 −1 0 (d) 20 r=1 r=10 r=100 r= ∞ σ=0.3 (c) 20 ) z H ( 0 ν 15 10 5 0 0 ) z H ( 0 ν 0.5 µ 1 15 10 5 0 0 2 10 1 r=1 r=10 r=100 r=1 r=10 r=100 r= ∞ µ=0 5 t(ms) 0.5 σ FIG. 1: (color online) (a) Illustration of the model. (b) V (t) trajectories for identical noise and three different values of r. (c) and (d) show the dependence of stationary firing rate on mean input current and noise strength in the noise driven regime. (τm = 10 ms, vr = 0, vb = 10, τr = 0) value ν1(ω) are the transmission functions and the phase angles φ1(ω) = arg(ν1(ω)) give the phase lags, which completely characterize the linear response. Note that we refer to both signal channels when the subscripts c and n are omitted. It is known that the absorbing boundary condition at vb can induce severe artifacts in the dynamical re- sponse. The potential vb marks a "point of no re- turn," which is not present in a biophysical dynamical model of AP initiation. As a consequence, the trans- mission function for a noise coded signal in the LIF model, for instance, does not decay at high signal fre- quencies [5, 11]. Ideally one would thus wish to sepa- rate the response function into a physiologically mean- ingful part νphy (ω) and a part containing all artifacts 1 such that νphy (ω). νphy (ω) = ν1(ω) − νabs (ω) must have 1 the following properties: i) νphy (ω) approaches the static susceptibility when ω → 0, specifically, νphy 1c (ω) → ∂ν0 and νphy (ω) → 0 when ω → ∞; iii) no essential dependence on the truncation point vb. The artifactual part from the absorbing boundary should have the following properties: i) negligible contribution for signal frequency in the physiologically relevant range f ≤ 1kHz, e.g. (ω), where f = ω/2π; ii) strong dependence on the truncation point vb. As we will show next such an isolation of the physiologically meaningful response is possible in the model Eq. (1). (ω) ≪ νphy 1n (ω) → 1 ∂σ ; ii) νphy νabs ∂ν0 ∂µ σ 1 1 1 1 1 1 In the model Eq. (1), the linear response can be ob- tained analytically by expanding the probability density in Eq. (2) to first order in ε and using the Green's func- tion method. We find that ν1n(ω) decomposes naturally into two parts, ν1(ω) = νLow (ω) + νHigh (ω), with 1 1 νLow 1c (ω) = νLow 1n (ω) = iω (1 − iω)(1 + iω/r) iω(iω − 1) (2 − iω)(2 + iω/r) (a) c 1 ν e d u t i l p m A 100 10−2 ν High→ 1c ← ν Low 1c (b) c φ g a L e s a h P 2 0 −2 100 102 f(Hz) (c) n 1 ν e d u t i l p m A 100 10−2 ν High→ 1n ← ν Low 1n 100 102 f(Hz) (d) n φ g a L e s a h P 100 2 0 −2 100 102 f(Hz) 102 f(Hz) FIG. 2: (color online) The normalized function ν1(ω)/ν1(0.1) and phase lag for a mean coded signal and noise coded signal with r = 1, µ = 0, and ν0 = 5 Hz. Other parameters are the same as in Fig. 1. Lines, theory; Symbols, simulation. where D = 1 2 σ2, ∆0 = (1 − vr)(2µ − 1 − vr)/4D and ∆1 = (1 − vb)(2µ − 1 + r(vb − vt))/4D. ψ1, Φ1, and Υ1 are parabolic cylinder functions, and Y1, Y2 are a combi- nation of parabolic cylinder functions, whose definition together with the νHigh (ω) parts are given in the sup- plement [14]. Any prime represents the derivative with respect to v. Here and in the following, the functions adopt their values at v = 1 if not denoted explicitly. Fig- ure 2 illustrates the linear response with r = 1 as an example. 1 Removing boundary induced artifacts. -- The decompo- sition of ν1(ω) into two additive components has exactly the features required for the separation of artifacts. Us- ing asymptotic expansion of the parabolic cylinder func- tions we find that for a finite signal frequency, νHigh can be approximated by 1 νHigh 1c (ω) ≃ νHigh 1n (ω) ≃ − ν0 r(vb − vt) + µ ν0 iω r + iω , (r(vb − vt) + µ)2 iω(1 + iω/r) 2 + iω/r . (4) 1 when vb ≫ vt. So νHigh (ω) are strongly dependent on vb and approach zero when vb → ∞ for finite signal frequency. When ω → 0, νHigh (ω) are negligible com- pared with νLow (ω) and are strongly suppressed when vb is large. That νHigh (ω) captures all artificial con- 1 1 1n 01) − (1 + iω/r)Φ1(vr)e∆0+iωτr , ′ (1 + 1/r)(ψ1P01 − √DΦ1P (1 + 1/r)(cid:0) ψ1(vr)e∆0+iωτr + (Y1ψ ′ Φ1P01 + 2Υ1P (1−iω)√D iω 1 ψ1)e∆1 ′ 1 − Y ′ 01(cid:1) + ν0 1 − Y ′ ψ1(vr)e∆0+iωτr + (Y1ψ ′ 1 ψ1)e∆1 D (2 + iω/r)Υ(vr)e∆0+iωτr 3 , (3) 1 1c 1n νLIF 1c (ω) → ν0 D = limω→∞ (ω) → ν0√D √ω eiπ/4 = limω→∞ tributions imposed by the absorbing boundary condi- tion is finally confirmed from the high frequency behav- ior, νHigh (ω) and νHigh νLIF 1n (ω), since the high fre- 1n quency behavior in the LIF model is determined solely by the absorbing boundary. As a consequence, neglect- ing νHigh (ω) in the response function eliminates any boundary induced instantaneous response components. These results establish that νLow (ω) capture the be- havior of ν1(ω) for low and intermediate frequencies and decay to zero in the large frequency limit. When ω → 0, νLow (ω) is negligible there. νLow (ω) exhibits only a weak depen- dence on vb through a frequency dependence phase lag φ0 = ω , characterizing the time lag to the trun- cation point vb of the AP upstroke. Therefore, we have νLow (ω). The physiologically meaningful pre- 1 dictions of the model can thus be revealed by examining νLow 1 1n (ω) → ν1n(0) = 1 ∂σ , since νHigh (ω) in isolation. r log vb+µ √rD (ω) = νphy ∂ν0 1n σ 1 1 1 1 Cutoff frequency and AP onset rapidness. -- Figure 3 shows the behavior of νLow (ω) with increasing r and how the cutoff frequency fc changes with r for different ν0 [15]. We see that fc increases linearly with the onset rapidness r for noise coded signals when the firing rate is not very low (> 1 Hz here); while for mean coded signals fc saturate for large r. The increase of fc with firing rate ν0 results from the stochastic double resonance phenomenon: the transmission function will develop a peak for some optimal signal frequency before decaying when ν0 is relatively large [16]. Figure 3D suggests that the cutoff frequency for a noise coded signal follows fc ∝ r and dissociates from τm. This is directly confirmed by the large frequency approximation of νLow 1n (ω), exp(− π 2 + iω/r (cid:18) − P 4 ω/r) ′ ′ (5) 01 + ipiω/r 2√D ν Low 1n (ω) ∝ 01(v0) → − ν0 P01(cid:19) , where P01(v) ≡ √rP01(v). Because P01(v0) → p π D for r ≫ 1, the decay of νLow ν0√D 1n (ω) and P depends essentially only on ω/r. This implies that the cutoff frequency for a noise coded signal dissociates from τm and becomes proportional to the onset rapidness r, fc = Ar, where A depends on ν0 through the effect dis- cussed above. This demonstrates that fast onset APs can enhance the cutoff frequency and therefore the response speed significantly. Note that a linear relationship fc ∝ r 2 (a) c 1 l w o ν e d u t i l p m A (c) n 1 l w o ν e d u t i l p m A 100 10−2 10−4 100 100 10−2 10−4 100 r=1 r=10 r=100 r=1000 r=∞ 102 f(Hz) r=1 r=10 r=100 r=1000 r=∞ 102 f(Hz) (b) (b) ) z H ( ) c 1 ν ( c f 300 250 200 150 100 50 104 0 50 r 100 (d) (d) ) z H ( ) n 1 ν ( c f 3000 2500 2000 1500 1000 500 104 0 50 r 100 1 FIG. 3: (color online) (a) and (c) The normalized transmis- sion function ν Low (ω)/ν1(0.1) for different r with µ = 0, ν0 = 5 Hz. (b) and (d) The variation of cutoff frequency with r for different firing rates: ν0 = 1, 5, 10, 20, 30, 40 Hz from lower to upper curves. Other parameters are the same as in Fig. 1. (a) 100 c 1 ν e d u t i l p m A 10−2 100 r=10 τ s τ s τ s τ s τ s =0 =1 =5 =10 =20 102 f(Hz) (b) 100 c 1 ν e d u t i l p m A 10−2 100 τ =0 s τ =1 s τ =5 s τ =10 s τ =20 s r=100 102 f(Hz) FIG. 4: Variation of the normalized transmission function ν1c(ω)/ν1c(0.1) with increasing correlation time τs (with unit ms) of the synaptic noise for r = 10, 100. Parameter used: µ = 0, vb = 10, ν0 = 5Hz. was previously conjectured by Naundorf et al. based on dimensional analysis [18]. 1 1c For a current coded signal, however, νLow (ω) ∝ √ω exp(− π 4 ω/r) for large r. Therefore the linear response is confined by the membrane time constant in the white noise case, as seen also from Fig. 3A. Real synaptic in- puts, however, have a finite correlation time and can be approximated better as a colored noise. As shown in Fig. 4, the confinement by the membrane time constant is broken under such conditions and cutoff frequencies of several hundred Hz can be reached for large r [14]. Our results identify AP onset rapidness as a critical de- terminant of population cutoff frequency and reveal how this cutoff frequency can dissociate from the basic single neuron time constants set by the mean firing rate and 4 membrane time constant. The confinement of the mean response for white noise constitutes an interesting predic- tion of the model, which should be tested experimentally by using a very short correlation time (≤ 1 ms) of the synaptic noise. The origin of the large onset rapidness seen in cortical neurons is a matter of ongoing debate [17, 18]. Its value can be modified in real neurons by applying drugs like TTX or knockout of sodium channel subtypes [12, 19]. Measurements of dynamical response for cortical neurons under such manipulations are thus predicted to provide important insight into the mecha- nism of fast population coding. We thank D. Battaglia and B. Lindner for suggestions and carefully reading an early version of the manuscript. We are grateful to T. Geisel, M. Gutnick, M. Huang, M. Monteforte, T. Moser and T. Tchumatchenko for discus- sions. This work was supported by BMBF (01GQ07113), GIF (906-17.1/2006), BCCN II (01GQ1005B), and DFG (SFD899). [1] H. Kondgen et al., Cereb. Cortex 18, 2086 (2008). [2] C. Boucsein et al., J. Neurosci. 29, 1006 (2009). [3] M. Higgs, W. Spain, J. Neurosci. 29, 1285 (2009). [4] T. Tchumatchenko et al., to be published. [5] G. Silberberg et al., J. Neurophysiol. 91, 704 (2004). [6] N. Fourcaud-Trocm´e et al., J. Neurosci. 23, 11628 (2003). [7] N. Fourcaud-Trocm´e, N. Brunel, J. Comput. Neurosci. 18, 311 (2005). [8] B. Naundorf, T. Geisel, F. Wolf, J. Comput. Neurosci. 18, 297 (2005). [9] M. Richardson, Phy. Rev. E 76, 021919 (2007). [10] N. Brunel et al., Phys Rev Lett 86, 2186 (2001). [11] B. Lindner, L. Schimansky-Geier, Phys Rev Lett 86, 2934 (2001). [12] B. Naundorf, F. Wolf, M. Volgushev, Nature 440, 1060 (2006). [13] L. Badel et al., J. Neurophysiol. 99, 656 (2008). [14] See supplemental material at http://link.aps.org/ sup- plemental/10.1103/PhysRevLett.106.088102. ∂D and saturates to a constant ν0 [15] For LIF model, ν1n(ω) starts at zero signal frequency with a value ∂ν0 D at large frequency. So in our model there will be a transition be- tween these two plateaus before the decaying of transmis- sion function, since in large r limit it reduces to the LIF model. To characterize the decaying property and avoid the transition region, we extract the cut-off frequencies at which the normalized transmission function decays to 1/√10, which is smaller than the value usually adopted. [16] H. Plesser and T Geisel, Phys. Rev. E 59, 7008 (1999). [17] D. McCormick, Y. Shu, Y. Yu, Nature 445:E1-E2(2007). [18] B. Naundorf, F. Wolf, M. Volgushev, Nature 445:E2-E3 (2007). [19] M. Royeck et al., J. Neurophysiol. 100, 2361 (2008).
1507.05970
1
1507
2015-07-21T20:05:42
Chaotic Neuronal Oscillations in Spontaneous Cortical-Subcortical Networks
[ "q-bio.NC" ]
Oscillatory activities are widely observed in specific frequency bands of recorded field potentials in different brain regions, and play critical roles in processing neural information. Understanding the structure of these oscillatory activities is essential for understanding the brain function. So far many details remain elusive about their rhythmic structures and how these oscillations are generated. We show that many oscillatory activities in spontaneous cortical-subcortical networks, such as delta, spindle, gamma, high-gamma and sharp wave ripple bands in different brain regions, are genuine chaotic time series which can be reconstructed as chaotic attractors through appropriately selected embedding delay and dimension. The reconstructed attractors are approximated by a simple radial basis function enabling high precision short-term prediction. Simultaneously recorded oscillatory activities in multiple brain regions differ greatly in term of temporal phase and amplitude but can be approximated by the same function. Our results suggest that neural oscillations are produced by deterministic chaotic systems. The occurrence of neural oscillation events is predetermined, and the brain possibly knows when and where the information will be processed and transferred in the future time as a result of the deterministic dynamic.
q-bio.NC
q-bio
Chaotic Neuronal Oscillations in Spontaneous Cortical- Subcortical Networks Pengsheng Zheng [email protected] [email protected] Abstract Oscillatory activities are widely observed in specific frequency bands of recorded field potentials in different brain regions, and play critical roles in processing neu- ral information. Understanding the structure of these oscillatory activities is es- sential for understanding the brain function. So far many details remain elusive about their rhythmic structures and how these oscillations are generated. We show that many oscillatory activities in spontaneous cortical-subcortical networks, such as delta, spindle, gamma, high-gamma and sharp wave ripple bands in different brain regions, are genuine chaotic time series which can be reconstructed as chaotic attractors through appropriately selected embedding delay and dimension. The re- constructed attractors are approximated by a simple radial basis function enabling high precision short-term prediction. Simultaneously recorded oscillatory activities in multiple brain regions differ greatly in term of temporal phase and amplitude but can be approximated by the same function. Our results suggest that neural os- cillations are produced by deterministic chaotic systems. The occurrence of neural oscillation events is predetermined, and the brain possibly knows when and where the information will be processed and transferred in the future time as a result of the deterministic dynamic. 1 Introduction Our brains are foretell machines with advanced predictive powers which are suggested to emerge from various neuronal oscillations (Buzsaki, 2006). These neuronal oscilla- tory activities prevail in cortical filed potentials in different behavior contexts demon- strating a number of characteristic oscillatory frequency bands, which are crucial for effective information processing within and across distinct brain structures (Buzsaki, 2006; Engel et al., 2001; Singer, 1999). Neural oscillations and synchronization have been linked to many cognitive functions such as sleep and consciousness, perception, motor coordination and memory (Fell and Axmacher, 2011; Schnitzler and Gross, 2005; Fries, 2005). Recent findings indicated that network oscillations temporally link neurons into as- semblies and facilitate synaptic plasticity (Buzsaki and Draguhn, 2004). The gamma cycle has been suggested to serve as a computational mechanism for the implementation of a temporal coding scheme that enabled fast processing and flexible routing of activity, supporting selection and binding of distributed responses (Fries et al., 2007). Power- frequency and spatial coherence analyses showed that gamma oscillations in hippocam- pal CA1 area split into distinct fast and slow frequency components routing the flow of in- formation (Colgin et al., 2009). Spontaneously occurring oscillatory events like thalamo- cortical sleep spindles and hippocampal sharp wave ripple complexes (SPW-R, 140- 250Hz) during slow wave sleep (SWS) were crucial for hippocampal-cortical interactions during memory consolidation (Wilson and Mcnaughton, 1994; Siapas and Wilson, 1998; 1 Sirota et al., 2003; Diba and Buzsaki, 2007; Eschenko et al., 2008; Girardeau et al., 2009; Logothetis et al., 2012). Understanding how these oscillations are generated and their rhythmic structures are fundamental and critical for the understanding of brain function. In this paper, we use two open access data sets, in one of which rats were implanted with silicon probes to monitor both LFP and unit firing in different hippocampal regions, such as CA1, entorhinal cortex layer 2 (EC2) and layer 5 (EC5) while they were sleep- ing (Mizuseki et al., 2014). In the other data set, the recordings were made from anterior thalamus (AT), hippocampus CA1 and mPFC of sleeping mice (Peyrache and Buzski, 2015). The simultaneously recorded field potentials in different regions teem with com- plex rhythmic structures which differ widely in term of phase, amplitude and temporal resolution. However, we show that these diverse rhythmic structures come from deter- ministic systems, and there is no random components in it. 2 Reconstruction of neuronal oscillatory activity We firstly band-pass filter the LFP to certain frequency range. Unless otherwise stated, we use rat CA1 spindle (12-15Hz) band activity during SWS (ec014.716.dat (Mizuseki et al., 2014)) as an example, and the results were extended to many other oscillatory activities (see Fig.S1-S12 in supplementary material (SM)). The spindle band activity gives us the time sequence of the form [X(t0), X(t1), · · · , X(t)], where X(t) denotes the oscillation voltage at time t (Fig.1A). Following the delay coordinate embedding method (Takens, 1981), the reconstructed attractor of the original system is given by the vector sequence P (t) = [X(t), X(t − τ ), X(t − 2τ ), · · · , X(t − (d − 1)τ )], (1) where τ and d are the embedding delay and embedding dimension respectively. The delay embedding theorem states that this procedure provides a one-to-one image of the original system for a large enough d (Takens, 1981). The embedding delay (τ ) is estimated by mutual information method which measures how much one knows about X(t−τ ) from X(t) (Fraser and Swinney, 1986). In principal, τ has to be large enough so that the information we get from measuring X(t − τ ) is significantly different from that of X(t), but τ should not be larger than the typical time in which the system loses information of its initial state. If τ is chosen larger than it should, then the reconstructed attractor will be folded and points will look more or less random since they will be uncorrelated. In practice, the τ can be estimated as the point where the average mutual information (in bits) is close to but before its first local minimum (τ = 0.015 in Fig.1B). The embedding dimension (d) is the minimum dimension of the space in which the trajectory does not statistically cross itself. As shown in Fig.1C, the embedding dimension is d = 5 suggested by Cao's method (Cao, 1997). We then reconstruct the 5-dimensional attractor and project it to 3-D (Fig.1D) and 2-D (Fig.1E) spaces, which are reminiscent of a chaotic attractor (compare Lorenz attractor Fig.S19 in SM). After reconstructing the attractor, we perform the determinism test to verify if the studied time series are indeed from a deterministic system (Kaplan and Glass, 1992). The determinism test constructs the vector field of the system directly from the time series, and subsequently tests if the reconstructed vector field assures uniqueness of solutions in the phase space (Kaplan and Glass, 1992), which enables us to distinguish between deterministic chaos and irregular random behavior that often resembles chaos. The calculated determinism measure for reconstructed spindle attractor is 0.99 (1 for perfectly deterministic), which means the system is deterministic. 2 2 4 6 10 Time (s) 8 B C ) d ( 1 E 0.8 0.6 0.4 0.2 12 14 16 18 20 0.015 0.03 0.045 τ 1 2 3 4 5 6 7 8 9 10 d 0.5 A V m 0 −0.5 0 t i B 6 4 2 0 D ) τ 2 − t ( X 0.5 0.4 0.3 0.2 0.1 0 −0.1 −0.2 −0.3 −0.4 −0.5 E ) τ 2 − t ( X 0.6 0.4 0.2 0 −0.2 −0.4 −0.6 −0.6 −0.4 −0.2 0 X(t−τ) 0.2 0.4 0.6 0.5 0 X(t) 0.5 0.4 0.3 0.2 0.1 0 −0.1 −0.2 −0.3 −0.4 −0.5 −0.5 X(t−τ) Figure 1: Reconstruction of spindle band activity. A: Time series of CA1 spindle. B: Average mutual information in dependence on embedding delay (τ ). C: E1(d), the Cao's-variation from d to d + 1, as a function of embedding dimension. D, E: Recon- structed attractor projected to 3-D and 2-D spaces respectively. 3 Neuronal oscillatory activity approximation and predic- tion Following the line of traditional method for chaotic time series prediction, thousands of input P (t) and output X(t + τ ) pairs are collected as t increases from time 0 to tp, and tp represents the ending time of training time series. It then reduces to a function approximation problem. Since the system is deterministic, it should be able to find a function to approximate all the input-output pairs. We here use Radial Basis Function Network (RBF) to approximate the delicate input-output function (Park and Sandberg, 1991). The RBF network has simple one-hidden layer structure and fast training process, and most importantly it has a more intuitive mathematical interpretation than the multilayer perceptron as follows, X(t + τ ) = LW · Radbas(kIW − P (t)k · b1) + b2, (2) where IW is input weight matrix, LW is weight matrix of output layer, b1 and b2 are bias terms of input and output layers respectively, k·k represents the Euclidean distance, and 2 Radbas(x) = ex . Here we use 'newrb' function in the Matlab neural network toolbox (Beale et al., 2013) to find proper parameters IW , LW , b1 and b2 for equation (2). The RBF network reaches a small goal of training error (MSE = 1e − 6) in a few iterations implying all the input-output pairs are well-fitted by the same network. The weight matrixes and bias terms of RBF network are frozen after training. 3 1 A C 0.3 0.2 0.1 0 -0.1 -0.2 -0.3 2 C E 0.1 0.05 0 -0.05 -0.1 5 C E 0.3 0.2 0.1 0 -0.1 -0.2 -0.3 A 0 0.5 1 1.5 MSE=8.76e-06 2 2.5 3 3.5 B 0 0.5 1 1.5 MSE=1.21e-06 2 2.5 3 3.5 C 0 0.5 1 1.5 MSE=6.24e-06 2 2.5 3 3.5 s Figure 2: Short-term prediction of spindle band activity. A-C, Comparison of predicted (red) and actual (blue) signals in CA1, EC2 and EC5 regions respectively. Note that all the predictions are performed by the network designed for CA1 spindle. MSE is the mean squared error. The prediction is applicable in real time, we are now able to predict the future value X(t + τ ), t > tp, by using P (t) and the frozen equation (2). For short-term prediction, P (t) is always from real signal, but none of these data is exposed to the training of RBF network. Figure 2A shows an example of the predicted X(t + τ ), the real signal (blue curve) disappears under the predicted one (red curve) as a result of high-accuracy prediction (MSE = 1.95e − 8). Many other spindle band activities collected from the same rat at different time were also tested. The network learned the data structure from a piece of signal and generalized to unknown signals. In other words, the spindle band oscillatory activity, no matter in the past or future, can be fitted into the same network (i.e. frozen equation (2)), which again proves the spindle activity is dominated by a deterministic system, and there is no irregular or random components in it. With the aforementioned method, we further verify that EC2 and EC5 spindle band activities are all developed by a deterministic chaotic system with the same embed- ding delay and dimension as that of the CA1 region. More interestingly, the EC2 and EC5 spindle band activities are accurately predicted by the RBF network intently de- signed for CA1 spindle time series (Fig.2B-C). Similarly, an RBF network trained by EC spindle band time series is capable of predicting CA1 spindle band activity. These all suggest that spindle band activities in different regions are dominated by deterministic dynamical systems which share similar or even identical structure. How could we reconcile the tremendous pointwise differences of simultaneously recorded multi-region signals with the facts that they were approximated by the same function? Take two identical Lorenz systems for example, the difference in initial states leads to vastly different trajectories (See Fig.S20 in SM). This implies the dynamical systems 4 v m 0.2 0.15 0.1 0.05 0 −0.05 −0.1 −0.15 −0.2 0.2 0.15 0.1 0.05 0 −0.05 −0.1 −0.15 −0.2 −0.25 0.05 0.1 0.15 t(s) 0.2 0.25 0.3 0.05 0.1 0.15 t(s) 0.2 0.25 0.3 0.2 0.15 0.1 0.05 0 −0.05 −0.1 −0.15 −0.2 0.2 0.15 0.1 0.05 0 −0.05 −0.1 −0.15 0.05 0.1 0.15 t(s) 0.2 0.25 0.3 0.05 0.1 0.15 t(s) 0.2 0.25 0.3 Figure 3: Examples of long-term iterative prediction of CA1 spindle activity. Red and blue curves represent the predicted and real signals respectively. Note that none of these data is exposed to the training of the network. in different regions might be identical in structure but operated with different initial states. Same argument, Gamma band activities in these three regions have the same em- bedding delay and dimension, and can be approximated by the same network intently designed for one region. The results can also extend to Delta, High-Gamma and SPW-R band activities with different dynamical systems developed for different frequency bands (see Fig.S1-S12 in the SM). In long-term prediction, the predicted value X(t + τ ) can be further applied to the prediction of X(t + 2τ ), and so forth. Unlike the short-term prediction, P (t) (t > tp + τ ) now is composed of predicted values rather than real signals. Figure 3 shows four examples of long-term iterative prediction of CA1 spindle band activity by the same frozen network. In all cases, the network achieves high-precision prediction at the beginning, and the performances extend for a certain amount of time followed by decreasing precision as time elapses. This is consistent with the observation of typical chaotic systems in which long term prediction is only theoretically plausible within a limited period of time due to its exponential divergence nature. All the results are further verified on the open access data set recorded in Hippocampal- mPFC-thalamic network of sleeping mice (Peyrache and Buzski, 2015), an RBF network is designed with the time series from one brain region and then applied to the prediction of signals in other brain regions. Figure 4 shows the results of short-term prediction of spindle band activity. More technical details and the results of other frequency bands are reported in Fig.S16-S17 (SM). Unless otherwise stated, all the results are retrieved during SWS state, but the network designed during SWS state is capable of predicting the signal in REM state (Fig.S18 SM). Matlab code is provided in the SM for easy and fast test of users' own neural signal under different conditions. 5 C P H 0.2 0.1 0 −0.1 −0.2 0 C F P m 0.1 0.05 0 −0.05 −0.1 0 T A 0.2 0.1 0 −0.1 −0.2 0 0.5 0.5 0.5 1 1 1 1.5 1.5 1.5 2 MSE=1.50e−06 2 MSE=3.53e−07 2 MSE=7.18e−07 2.5 2.5 2.5 3 3 3 3.5 3.5 3.5 4 4 4 (s) Figure 4: Short-term prediction of spindle band activity in sleeping mouse (Mouse12- 120807). Comparison of predicted (red) and actual (blue) time series in hippocampal (HPC) CA1, mPFC and AT regions respectively. Note that all the predictions are performed by the network designed for HPC CA1 spindle. 4 Conclusion and discussion Many cognitive functions of the brain are suggested to be emerged from neural oscil- lations, but where do these complex oscillations come from? Here we show that these diverse rhythmic activities, in different brain regions but in the same frequency band, come from a unified deterministic system, and there is no random components in it. The occurrence of certain kinds of neural oscillation events is predetermined by the deter- ministic dynamic rather than randomly or statistically formed. In this sense, the brain has perfect control of local and global information flow, which is ideal for information transfer, avoidance and collaboration between distinct regions. There is a long history of chaotic system study in physics and many interdisciplinary areas with numerous proposed theorems and applications. Now the whole theory is potentially open to neural oscillation research, and many interdisciplinary topics are expected to be explored in the future. On the other hand, neural oscillations have been suggested to play critical roles in many cognitive functions such as sleep, conscious- ness, perception and memory. The results of paper provide potential new perspectives to many of these researches. For example, among all the topics of chaos theory, the synchronization of multiple chaotic attractors is one of the hottest topics. As men- tioned above, two identical Lorenz systems display vastly different trajectories when they are operated at different initial states (Fig.S21 in SM). However, the two systems can be fast synchronized if a negative feedback is applied (Fig.S15 in SM). The syn- chronization level is further adjustable by tuning the strength of the negative feedback which might be analogue to the changes of network connectivity wildly observed in epilepsy (David et al., 2008; Zhang et al., 2011; Clemens et al., 2013). Integration of the results with the epilepsy research might be one of the interesting studies in the future. 6 References Beale, M., Demuth, H., and Hagan, M. (2013). Neural Network Toolbox User's Guide. The MathWorks Inc. Buzsaki, G. (2006). Rhythms of the brain. Oxford University Press. Buzsaki, G. and Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science, 304(5679):1926 -- 1929. Cao, L. Y. (1997). Practical method for determining the minimum embedding dimension of a scalar time series. Physica D, 110(1-2):43 -- 50. Clemens, B., Puskas, S., Besenyei, M., Spisak, T., Opposits, G., Hollody, K., Fogarasi, A., Fekete, I., and Emri, M. (2013). Neurophysiology of juvenile myoclonic epilepsy: EEG-based network and graph analysis of the interictal and immediate preictal states. Epilepsy Research, 106(3):357 -- 369. Colgin, L. L., Denninger, T., Fyhn, M., Hafting, T., Bonnevie, T., Jensen, O., Moser, M. B., and Moser, E. I. (2009). Frequency of gamma oscillations routes flow of information in the hippocampus. Nature, 462(7271):353 -- U119. David, O., Guillemain, I., Saillet, S., Reyt, S., Deransart, C., Segebarth, C., and De- paulis, A. (2008). Identifying neural drivers with functional MRI: An electrophysio- logical validation. Plos Biology, 6(12):2683 -- 2697. Diba, K. and Buzsaki, G. (2007). Forward and reverse hippocampal place-cell sequences during ripples. Nature Neuroscience, 10(10):1241 -- 1242. Engel, A. K., Fries, P., and Singer, W. (2001). Dynamic predictions: Oscillations and synchrony in top-down processing. Nature Reviews Neuroscience, 2(10):704 -- 716. Eschenko, O., Ramadan, W., Molle, M., Born, J., and Sara, S. J. (2008). Sustained in- crease in hippocampal sharp-wave ripple activity during slow-wave sleep after learning. Learning & Memory, 15(4):222 -- 228. Fell, J. and Axmacher, N. (2011). The role of phase synchronization in memory pro- cesses. Nature Reviews Neuroscience, 12(2):105 -- 118. Fraser, A. M. and Swinney, H. L. (1986). Independent coordinates for strange attractors from mutual information. Physical Review A, 33(2):1134 -- 1140. Fries, P. (2005). A mechanism for cognitive dynamics: neuronal communication through neuronal coherence. Trends in Cognitive Sciences, 9(10):474 -- 480. Fries, P., Nikolic, D., and Singer, W. (2007). The gamma cycle. Trends in Neurosciences, 30(7):309 -- 316. Girardeau, G., Benchenane, K., Wiener, S. I., Buzsaki, G., and Zugaro, M. B. (2009). Selective suppression of hippocampal ripples impairs spatial memory. Nature Neuro- science, 12(10):1222 -- 1223. Kaplan, D. T. and Glass, L. (1992). Direct test for determinism in a time-series. Physical Review Letters, 68(4):427 -- 430. 7 Logothetis, N. K., Eschenko, O., Murayama, Y., Augath, M., Steudel, T., Evrard, H. C., Besserve, M., and Oeltermann, A. (2012). Hippocampal-cortical interaction during periods of subcortical silence. Nature, 491(7425):547 -- 553. Mizuseki, K., Sirota, A., Pastalkova, E., Diba, K., Peyrache, A., and Buzski, G. (2014). Extracellular recordings from multi-site silicon probes used for cluster- ing of neuron responses in rat hippocampal and entorhinal regions. CRCNS.org, http://dx.doi.org/10.6080/K0QJ7F75. Park, J. and Sandberg, I. W. (1991). Universal approximation using radial-basis-function networks. Neural Computation, 3(2):246 -- 257. Peyrache, A. and Buzski, G. (2015). Extracellular recordings from multi-site silicon probes in the anterior thalamus and subicular formation of freely moving mice. CR- CNS.org, http://dx.doi.org/10.6080/K0G15XS1. Schnitzler, A. and Gross, J. (2005). Normal and pathological oscillatory communication in the brain. Nature Reviews Neuroscience, 6(4):285 -- 296. Siapas, A. G. and Wilson, M. A. (1998). Coordinated interactions between hippocampal ripples and cortical spindles during slow-wave sleep. Neuron, 21(5):1123 -- 1128. Singer, W. (1999). Neuronal synchrony: A versatile code for the definition of relations? Neuron, 24(1):49 -- 65. Sirota, A., Csicsvari, J., Buhl, D., and Buzsaki, G. (2003). Communication between neocortex and hippocampus during sleep in rodents. Proceedings of the National Academy of Sciences of the United States of America, 100(4):2065 -- 2069. Takens, F. (1981). Detecting strange attractors in turbulence. Lecture Notes in Mathe- matics, 898:366 -- 381. Wilson, M. A. and Mcnaughton, B. L. (1994). Reactivation of hippocampal ensemble memories during sleep. Science, 265(5172):676 -- 679. Zhang, Z. J., Valiante, T. A., and Carlen, P. L. (2011). Transition to seizure: From "macro"- to "micro"-mysteries. Epilepsy Research, 97(3):290 -- 299. 8
1804.00153
2
1804
2018-04-06T09:06:53
Spontaneous activity emerging from an inferred network model captures complex temporal dynamics of spiking data
[ "q-bio.NC" ]
The combination of new recording techniques in neuroscience and powerful inference methods recently held the promise to recover useful effective models, at the single neuron or network level, directly from observed data. The value of a model of course should critically depend on its ability to reproduce the dynamical behavior of the modeled system; however, few attempts have been made to inquire into the dynamics of inferred models in neuroscience, and none, to our knowledge, at the network level. Here we introduce a principled modification of a widely used generalized linear model (GLM), and learn its structural and dynamic parameters from ex-vivo spiking data. We show that the new model is able to capture the most prominent features of the highly non-stationary and non-linear dynamics displayed by the biological network, where the reference GLM largely fails. Two ingredients turn out to be key for success. The first one is a bounded transfer function that makes the single neuron able to respond to its input in a saturating fashion; beyond its biological plausibility such property, by limiting the capacity of the neuron to transfer information, makes the coding more robust in the face of the highly variable network activity, and noise. The second ingredient is a super-Poisson spikes generative probabilistic mechanism; this feature, that accounts for the fact that observations largely undersample the network, allows the model neuron to more flexibly incorporate the observed activity fluctuations. Taken together, the two ingredients, without increasing complexity, allow the model to capture the key dynamic elements. When left free to generate its spontaneous activity, the inferred model proved able to reproduce not only the non-stationary population dynamics of the network, but also part of the fine-grained structure of the dynamics at the single neuron level.
q-bio.NC
q-bio
Spontaneous activity emerging from an inferred network model captures complex temporal dynamics of spiking data Cristiano Capone*,1, 2, a) Guido Gigante*,1, 3 and Paolo Del Giudice1, 2 1)Istituto Superiore di Sanit`a, Rome, Italy 2)INFN, Sezione di Roma 1, Italy 3)Mperience s.r.l., Rome, Italy The combination of new recording techniques in neuroscience and powerful inference methods recently held the promise to recover useful effective models, at the single neuron or network level, directly from observed data. The value of a model of course should critically depend on its ability to reproduce the dynamical behavior of the modeled system; however, few attempts have been made to inquire into the dynamics of inferred models in neuroscience, and none, to our knowledge, at the network level. Here we introduce a principled modification of a widely used generalized linear model (GLM), and learn its structural and dynamic parameters from ex-vivo spiking data. We show that the new model is able to capture the most prominent features of the highly non-stationary and non-linear dynamics displayed by the biological network, where the reference GLM largely fails. Two ingredients turn out to be key for success. The first one is a bounded transfer function that makes the single neuron able to respond to its input in a saturating fashion; beyond its biological plausibility such property, by limiting the capacity of the neuron to transfer information, makes the coding more robust in the face of the highly variable network activity, and noise. The second ingredient is a super-Poisson spikes generative probabilistic mechanism; this feature, that accounts for the fact that observations largely undersample the network, allows the model neuron to more flexibly incorporate the observed activity fluctuations. Taken together, the two ingredients, without increasing complexity, allow the model to capture the key dynamic elements. When left free to generate its spontaneous activity, the inferred model proved able to reproduce not only the non-stationary population dynamics of the network, but also part of the fine-grained structure of the dynamics at the single neuron level. Dynamic models, in neuroscience and in general, embody in a mathematical form the causal relationships between variables deemed essential to describe the system of interest; of course, the value of the model is measured both by its ability to match observations, and by its predictive power. Typically, along this route, parameters appearing in the model are assigned through a mix of insight from experiments and trial-and-error and, in turn, the study of the model dynamics helps understanding the relevance of each parameter in determining the dynamic regimes accessible to the system. Another approach, initially quite detached from dynamic modeling, is rooted in the domain of statistical inference, and a seminal example in neuroscience was offered by the application of maximum-entropy inference of Ising-like models to multi-electrode recordings of neural activity; in this case, as well as in the later extension to kinetic Ising-like models, the Montecarlo/Glauber dynamics of the model is only meant as a means to sample the probability distribution of interest and it is not claimed to offer a detailed model of the actual dynamics at work in the system. Recently, Generalized Linear Models (GLM) (which incorporate kinetic Ising models as a special case) have been recognized as flexible and powerful inference models1,2. In time, efforts have been made to make contact between the two approaches, e.g., in the case of neuroscience, by endowing the coupling structure of the inference model with features motivated by biological plausibility3–5; it has also been recognized that a GLM is close to a stochastic version of the spike-response model. Recently, the repertoire of the driven a)Electronic mail: [email protected] *authors contributed equally 2 dynamics of GLM models of single neurons has been explored6; however, to our knowledge, a largely open issue is to endow inference models with predictive power in terms of the system dynamics. Our approach to this problem is to explore the free, spontaneous dynamics of the inferred model in its relation with the one of the biological system generating the data and, in doing this, to identify the role of different elements of the inference model in determining the spontaneous dynamics of the neuronal network. Indeed, an obvious but persisting problem in the application of inference models to neuroscience has been how to assess the meaning and value of the inferred parameters; a recurring example is offered by the inferred synaptic couplings3: in the literature, a cautionary remark is usually included, recognizing that the inferred couplings (whether in the form of synaptic efficacies of more complicated synaptic kernels) are to be meant as 'effective', leaving of course open the problem of what exactly this means (especially in the face of the dramatic subsampling of the underlying biological network). The value of the inferred model cannot be assessed directly by a detailed correspondence between its elements and corresponding elements of the biological system; using the output of a GLM to decode input stimuli is an interesting recent approach7 Our attitude is that spontaneous activity is a good testing ground to assess whether the inferred model does indeed capture essential features of the biological system, and we choose a case in which the spontaneous activity of the biological network is highly non-stationary and irregular: a cultured neuronal network generating spontaneously a wide spectrum of activity fluctuations, from population bursts, to neuronal avalanches and noisy oscillations8–20. Such richness is extremely hard to reproduce with a GLM, and to our knowledge no models proposed so far were able to cope with it. For this reason we chose population bursting activity as a challenging benchmark to test our model. In our GLM approach we introduced novel ingredients inspired by biological observation, such as activity dependent negative feedback over different time-scales for the single neuron8 (spike frequency adaptation, SFA), a bounded transfer function, and a generative probabilistic mechanism for spike generation with super-Poisson fluctuations. We show that these ingredients are crucial to endow the system with non-stationary spontaneous activity and also to account for detailed dynamic features such as temporal correlations and spatio-temporal evolution of network bursts. RESULTS Bounded firing rate, super-Poisson spike generation, and their impact on the spontaneous activity of the inferred model The model we introduce is an extension of GLMs as widely employed for inferring the structure of neuronal networks. With GLMs it shares the three fundamental assumptions that spike generation is a probabilistic process driven by the instantaneous current afferent to the neuron; that, secondly, such current is a linear function of the activity of the neurons in the network and, finally, that there is a non-linear transfer function linking the current to the probabilistic spike generation. The parameters of the model, in this framework, are usually obtained through the numerical maximization of the likelihood that the model itself generated the observed data. The main points of departure between more standard GLMs and the present model lay in the choice of the specific non-linear transfer function and probabilistic generative model; another notable difference is in the introduction, at the single neuron level, of an activity-dependent self-inhibition current, mimicking spike- frequency adaptation (SFA) effects widely observed in real neurons. Fig. 1, panel A, gives a schematic view of the model. The current Hi(t) instantaneously felt by neuron i sums spikes emitted by other neurons j in the recent past (t− ∆t) through synaptic kernels kij(∆t), a constant external field hi, and a set of 5 SFA signals (collectively indicated as ci(t)), that integrate, over different time-scales, the spiking activity Si of neuron i itself. This current provides input to a non-linear transfer-function f [Hi], which in turn determines the expected firing rate λi(t) for the spike generation model at the next time bin t + dt. The transfer function and generative model most commonly employed in the 3 FIG. 1: Spontaneous activity of the inferred model reproduces non-stationary network dynamics. A Schematic representation of the model. The current Hi(t) felt by neuron i sums spikes emitted by other neurons j in the recent past (t − ∆t) through synaptic kernels kij(∆t), a constant external field hi, and spike-frequency adaptations signals that integrate, over different time-scales, the spiking activity Si of neuron i itself. This current, fed into a non-linear transfer-function f [Hi], determines the expected number λi(t) of spikes Si(t + dt) that the neuron will probabilistically generate at the next time bin. B Rastergrams of ex-vivo cultured cortical neurons show a strongly non-stationary dynamics (top row; red line: whole network activity). The proposed Sig-NegBin model (central row, blue line) closely follows the data used to drive its dynamics (red line, t between 20 and 30 s); when left free to run autonomously at t = 30 s, the model displays large irregular bursts of activity, qualitatively mimicking the behavior observed in the recordings. The standard Exp-Poisson model (bottom row, blue line), on the other hand, whilst still able to follow the data in driven mode, fails to produce any large activity fluctuation, settling instead into a noisy, stationary state. C Comparison between burst amplitudes (total spike counts, left), bursts durations (center) and IBI durations (right) histograms both for the data (top row) and the autonomous activity of the Sig-NegBin model (bottom row). Mean and standard deviation of each distribution are reported with red and green lines respectively. Results for the standard Exp-Poisson model are not shown because of the extreme sparsity of burst events in that case. D Average area under the inferred synaptic kernels (absolute value) as a function of the distance separating the electrodes; blue and red circles are for excitatory and inhibitory kernels respectively; dashed lines: exponential fits; the strength (area under the kernel) of inferred excitatory synapses decays faster than that of the inhibitory ones. 4 literature are, respectively, the exponential function and the Poisson distribution; in the following we will label this choice 'Exp-Poisson'; our model, on the other hand, makes use of a sigmoid function (see Materials and Methods) and of a Negative-Binomial distribution with shape parameter rNB set to 0.2, and will be referred to in the following as 'Sig-NegBin' model. The synaptic kernels kij(∆t), following4, are parametrized as a weighted sum of 4 'raised cosine' functions peaking at times spanning a range between 10 ms (equal to the chosen time bin dt) and 30 ms, with support extending in the past for 15 dt = 150 ms (kij(∆t) ≡ 0 for ∆t > 150 ms). The sigmoid transfer function has two free parameters: the maximum firing rate attainable by the neuron and an asymmetry parameter that determines how fast this maximum is approached for high currents compared to the other asymptote at 0 for strongly negative currents. Finally each of the 5 SFA signals ci, that are set equal for each neuron, has associated a characteristic time-scale τc for the integration of the spike train Si(t) and a weight gc with which ci(t) enters the sum of the current Hi(t). These latter parameters, the weights defining each of the kij(∆t) functions, and the free parameters of the sigmoid function are all determined through the likelihood maximization procedure on data (see Materials and Methods for further details). In the case of the Exp-Poisson model, the only optimized parameters are the kij(∆t) weights, both for i (cid:54)= j and i = j ('post-spike filter'); the kii functions are parametrized as a weighted sum of 10 raised cosines peaking at times spanning from 10 ms to 50 ms and again extending to 150 ms in the past; the Sig- NegBin model does not include post-spike filters; thus the total number of free parameters for the Sig-NegBin is actually lower than that for the Exp-Poisson model. Once the parameters are fitted by maximizing the likelihood, both models can be simu- lated to generate new spike trains according to two basic simulation modalities. The first one ('driven' mode) feeds the model with the actual spike trains observed in the data and at each time t uses the model's output to make a 1-step prediction of the network activity at step t + dt; thus the generated activity never re-enters the network dynamics; the prob- abilistic distance between the 1-step prediction and the actual recorded network activity is ultimately what is minimized during the optimization procedure. In the second modality ('free' mode) the models are left free to evolve and the generated spike trains are fed back into the network, driving its dynamics at subsequent times. We trained both the Sig-NegBin and the Exp-Poisson model on data from multi-electrode recordings of ex-vivo cultured cortical neurons10, that show a clear non-stationary dynamics (Fig. 1B, top row) in the rastergrams and in the whole network activity (red line), with large and irregular bursts of synchronous activity interleaved by periods of low, noisy activity ('inter-burst intervals' or IBIs). At the end of the training, both models show, in driven mode, an activity that follows quite closely the experimental spike trains (Fig. 1B, middle and bottom rows; red line: data; blue line and rastergrams: simulation). However, when the simulation is switched to free mode at time t = 30 s, the behavior of the two models clearly diverges: whilst the Sig-NegBin is able to autonomously sustain a highly non-stationary activity, the Exp-Poisson only exhibits stationary fluctuations. Note that, in free mode, the activity of the Sig-NegBin model does not closely follow the activity of the real network anymore (top row, last 10 s); this is not surprising: the choice of a probabilistic model as a GLM stems from the assumption that the network dynamics is, to a large extent, influenced by self-induced random fluctuations and, thus, non-deterministic; it is therefore expected that, at least after a transient phase, the model and the data will drift away. Yet the activity spontaneously generated by the model clearly resembles the one seen in data in an intuitive statistical way, showing irregular bursts interspersed with interval of relative quiescence. Such resemblance is made more quantitative in Fig. 1C. Bursts of the network global activity are detected through the algorithm developed in8, both for the data (top row) and the autonomous activity of the Sig-NegBin model (bottom row). The comparison between burst amplitudes (total spike counts, left), bursts durations (center) and IBI durations (right) histograms shows a semi-quantitative agreement between the dynamics of the bio- logical network and the inferred model (mean and standard deviation of each distribution 5 FIG. 2: Bounded spike rate is necessary to generate non-stationary spontaneous activity in the inferred model. A Population activity (blue line) and rastergram generated by the model where the bounded transfer function is replaced by an exponential and inference procedure is carried out from scratch. Red line represent the data driving the dynamics for the first 10 s (20 ≤ t ≤ 30 s); afterwards, when left free to generate its spontaneous activity, the model is unable to sustain any non-stationary bursting activity and settles in a quiescent state. B-C transfer function (respectively sigmoid and exponential) f [·] (top-right panel) distribution of input currents Hi(t) to the single neuron across different times (bottom panel) and distribution of firing rates λi(t) = f [Hi(t)] ( top-left panel) are reported with red and green lines respectively). The main discrepancy between model and data is visible in the IBI distribution; "doublets" of bursts in the data, clearly separated by the rest of the IBI distribution, are absent in the model. Based on previous modeling work on similar data21, the discrepancy is likely due to elements, we did not include in our minimal model, such as short-term synaptic facilitation and depression. We remark that the analogous distributions for the Exp-Poisson model are not reported because of the extreme sparsity of burst events in that case. Moreover we observe that, despite the major difference in the exhibited dynamic behavior, the inferred synaptic kernels for our model and the Exp-Poisson model are correlated (cPearson = 0.49 with a P < 0.05; correlation between the total areas under each synaptic kernel, as a measure of 'synaptic strength'). The inferred synaptic kernels also show a correlation with the distance separating the electrodes where the activity of the pre- and post-synaptic neurons where recorded (Fig. 1D; blue and red circles are for excitatory and inhibitory kernels respectively; dashed lines: exponential fits). It is worth noting that the connections' decay with distance is faster for excitatory than for inhibitory inferred synapses, consistently with what is found in activity correlations in the neocortex by using 2D multielectrode arrays22. The choice of a sigmoidal, saturating transfer function, besides seeming just natural in a GLM meant to model neural activities, turns out to be key in achieving the above results. We repeated the inference procedure with a hybrid Sig-NegBin model, in which the sigmoidal transfer function is replaced with an exponential (Fig. 2A). As above, when during the first 6 10 s the model is simulated (blue line and rastergrams) in driven mode, it shows a good agreement with the driving data (red line); when the driving is turned off, instead, the model settles in a quiescent state, absent any large bursts of activity. In free mode, the saturation of the sigmoidal function f [·] (Fig. 2B, top-right panel) allows a wide distribution of input currents Hi(t) to the single neuron across different times (Fig. 2B, bottom panel) to give rise to a distribution of firing rates λi(t) = f [Hi(t)] (Fig. 2B, top-left panel) that is strongly bimodal (notice the logarithmic scale on the y-axis). On the other hand, the fast growth of the exponential function (Fig. 2C, top-right panel) pushes the inference procedure towards somewhat narrower, more conservative distribution of currents (Fig. 2C, bottom panel), resulting in a unimodal distribution of firing rates. The comparison of the two models in driven and free modes suggests that the exponential transfer function makes the system very selectively susceptible to the fluctuations experienced during learning, while the sigmoid transfer function allows for susceptibility to a wider range of fluctuations, which we believe is the root of the ability to spontaneously generate network bursts. The recorded neurons from the cultured network constitute of course a dramatic sub- sampling of the network (Fig. 3A, leftmost sketch; though, in fact, the activity of several neurons will be collected by each electrode). The effects of the unobserved neurons on the behavior of the observed part of the network could be in principle very complex23; yet we argue that the adoption of a Negative Binomial distribution for the spike counts generation, instead of a Poisson distribution as commonly assumed, captures a relevant part of those effects and is in fact critical to account for the non-stationarity of the spontaneous activity of the network. (cid:80) One way to approach the problem is to assume that at each time t the generated spike count of a generic neuron Si(t) is determined by both the field Hi(t) = hi + j αij(τ ) Sj(t − τ ) (the sum of external fields and the one due to the activity of the other observed neurons), and the field due to inputs from unobserved, 'hidden' neu- rons, hhid (Fig. 3A), such that the observed spike count probability distribution would be obtained by integrating over the (unknown) distribution of hhid (cid:80) : τ i i (cid:90) p(SiHi) = i p(SiHi, hhid dhhid i ) p(hhid i ) (1) by λi(t) = f(cid:2)Hi(t) + hhid (cid:3). i where we still assume that p(Si(t)Hi(t), hhid) is a Poisson distribution, with mean given It is intuitive that p(Si(t)Hi(t)) will now generate super-Poissonian fluctuations; and it turns out that larger fluctuations are critical for the ability of the Sig-NegBin model to spontaneously generate bursting activity, as shown above (Fig. 3B upper panel; blue line: data; red line: Sig-NegBin model in free mode). In fact, an inferred Sig-Poisson model shows very sparse bursting (Fig. 3B, middle panel). But it is not just that a broader spike generation distribution favors larger fluctuations in the network activity. We simulated the Sig-NegBin model replacing the Negative Binomial with a Poisson spike generator (Fig. 3B, bottom panel); such 'Poisson replay' of the inferred Sig-NegBin model generates frequent bursts, each lasting for long time on average. Thus the inferred synaptic couplings of the Sig-NegBin model, unlike the ones for the Sig-Poisson, appear to be compatible with a bistable network dynamics, between UP and DOWN states, as also suggested by the results for the sigmoid transfer function (Fig. 2B). Therefore, the main role of a super-Poisson spike generation distribution seems to be to destabilize the UP state, making the inferred model once again more robust to the variability of the free network's activity and thus able to self-sustain a highly non-stationary activity. Taking together the results in Fig. 2 and Fig. 3, the suggested picture is that on the one hand the sigmoid gain function allows the inference procedure to explore and use ranges of couplings that are effectively forbidden for an exponential gain function, and are essential to spontaneously generate large sudden increase of activity (burst onset); on the other hand, the NegBin generator plays the dynamic role of efficiently destabilizing, with large fluctuations of the spike counts, the high-activity states which are spontaneously generated by the large recurrent excitation. The two ingredients are robustly coupled to spontaneously 7 FIG. 3: Rationale and implications of choosing a negative binomial as generative model of spike counts. A Sketch showing the effects of the unobserved neurons on the fluctuations in the observed part of the network. B Spontaneous bursting activity generated by different models (blue line: data; red line: model in free mode). Upper panel: Sig-NegBin model; middle panel: Sig-Poisson; bottom panel: Sig-NegBin model after replacing the Negative Binomial with a Poisson spike generator. C Comparison between the spike counts generated by a negative binomial distribution (r = 0.22) with a mean value equal to the average spike count of the neuron (blue line), the Monte Carlo sampled p(SiHi) with best matching values for the first two moments of spike counts (red line) and Poisson distribution for the same average spike count (black line). produce bursting activity. Albeit representing just one of the forms that p(Si(t)Hi(t)) can take, the NegBin distri- bution has been suggested to adequately model fluctuations in observed neural activity24,25, which makes it a natural candidate. Yet, a NegBin would result from Eq. 1 assuming an exponential transfer function and a log-Gamma distribution for hhid : the first condition is patently in contradiction with our choice of a saturating transfer function and the second one is not expected to hold in general. We nevertheless tested the adequacy of the NegBin assumption for the Sig-NegBin model as follows. We assumed that p(hhid ) in Eq. 1 is a scaled and shifted version of the distribution of the observed fields H; in turn we estimated such distribution from the inferred synaptic couplings and the observed experimental spike counts, for the neuron with maximal average activity, for which we expected the differences in the spike count distribution would both show up more clearly, and matter more for the dynamics. From this, we performed a Monte Carlo estimate of p(SiHi) for different values i i i of the mean and the standard deviation defining p(hhid ). Fig. 3C shows a NegBin that is very close to the one used in the Sig-NegBin model (r = 0.22) for a mean value equal to the average spike counts of the neuron (blue line); the sampled p(SiHi) with best matching values for the first two moments is reported in red; the two distributions are very similar, especially if compared to the Poisson distribution for the same average spike count (black line). Thus spike counts compatible with a NegBin distribution are naturally accounted for by a plausible assumption on the effect of sub-sampling underlying Eq. 1, with marked deviations from a Poisson distribution. We repeated the above procedure sampling hhid from a Gaussian distribution, finding similar results; this suggests that, provided suitable values for the mean and variance are chosen, the precise shape of the p(hhid ) has a minor effect. i 8 i The inferred model captures detailed temporal and spatial aspects of the neural dynamics Although the spike-frequency mechanism with which the Sig-NegBin model is endowed allows for 5 independent time-scales, it is interesting to note that the inferred values consis- tently aggregate around just two values, at about 100 ms and 2 s respectively, for different temporal segments of the same network's activity (Fig. 4A). This result is consistent with previous work on the same data aiming to infer the main time-scales of the bursting dy- namics via a completely different approach8. To assess how relevant the adaptation mechanism is for the model dynamics, we performed inference with and without SFA. Then we simulated the model in driven mode until just after the end of a burst of large amplitude, where adaptation effects are expected to be more pronounced, leaving the model free to spontaneously evolve afterwards. In Fig.4B the SFA and non-SFA dynamics are reported (top and bottom, respectively); black line is the model activity averaged over 500 simulations of the dynamics, while gray shading marks the plus/minus one standard deviation range; red line is the population activity from data. For each realization we collected the next IBI, that is the time interval to the next burst spontaneously generated by the model. The IBI histogram (inset) attains a maximum very close to zero when SFA is not present; such maximum shifts toward higher values with SFA; the two distributions also have different averages (2.3 s vs 3.2 s) and coefficients of variation (0.98 vs 0.65). Thus SFA significantly increases the average interval to the next burst, reducing at the same time its variability. We also found in data, consistently with the results reported in19, a positive correlation between burst amplitude and the length of the previous IBI, an effect that could be possibly attributed to some adaptation mechanism, such as SFA. We estimated this correlation from simulations of the models inferred, with and without SFA, on different recordings; Fig. 4C compares the result of the two models with that found in their experimental counterparts. It is seen that the model without SFA is unable to recover correlations that are significantly different from zero, while the model with SFA produces correlations that are consistent with that measured in the data. We also addressed the ability of the model to account for the temporal structure and spatial organization of the network activity. We first considered the average order of activa- tion of different spatial locations within a burst (see Methods for details): time-rank is the index of the time bin when a neuron spikes first since the burst onset. By evaluating the average rank for both each neuron in the model and each electrode in the data, we found them strongly correlated (c = 0.9). In Fig. 5A we report the spatial distribution of average time-ranks for data (left) and simulations (right). Looking farther into the burst temporal structure, we asked to what extent the time-ranks in the model and in the data are comparable on a single burst basis. For each burst in the data, we selected and compare the simulated burst with the closest time-rank pattern. Fig 5B and C show two examples of such comparison. To provide a global comparison, capturing the correlation in time-ranks between data and simulations, we proceeded as follows: independently for each neuron, we shuffled the vector of its time-ranks across 9 FIG. 4: Effect of inferred time-scales and SFA currents. A Inferred τSFA for different recordings for the same preparation; the five inferred time-scales cluster around two values for all recordings, one about 100ms, the other around 2s. B Comparison between the model dynamics with and without SFA. In black, and gray-shading, are represented the average and variance over 500 simulations. Red line is the average activity from data during a large chosen burst. Insets: histograms of the time to the subsequent burst. C Correlation between the burst amplitude and the time since the preceding burst. all the simulated bursts, thereby destroying spatial correlations in the time-ranks, while preserving the average ranks (Fig. 5A, right). shuf f and dmin For each burst in the data we took the closest simulated burst and the closest shuffled simulated one, computed the corresponding distances (dmin sim ) and the distribution of their differences (Fig. 5D). It is seen that, for the large majority of bursts, the simulated burst is closer to the data burst than the surrogate. We therefore can conclude that inferred model captures most of the spatial development of neural activity at the single burst level. To inquire into the relationship between the inferred synaptic structure and the ensuing network dynamics, we also asked whether for a neuron with low average time-rank (early spiking), the efficacy of its outgoing synapses correlates with the time-ranks of its post- synaptic neurons. Fig. 5E shows indeed a high negative correlation between the efficacy of outgoing synapses and the time-rank of post-synaptic neurons (c = −0.8). DISCUSSION Several criteria can be identified to evaluate a model, yet the ability to reproduce the behavior of the system under analysis is certainly among the most relevant. We have seen that the addition of few computational ingredients can enable a probabilistic, generative network to autonomously sustain a rich, highly non-stationary dynamics, as observed in the experimental data employed to infer the model's parameters. In recent years, the field of statistical inference applied to neuronal dynamics has mainly focused on devising models and procedures that could reliably recover the real or effec- tive synaptic structure of a network of neurons, under different dynamical and stimulation conditions26–28, even though, more often than not, an experimental ground truth was not readily available. Although, quite surprisingly, the study of the dynamics of the inferred models has been largely neglected, we are not the first to address the issue. In4 the authors demonstrated, on an in-vitro population of ON and OFF parasol ganglion cells, the ability of a GLM to accurately reproduce the dynamics of the network;7 studied the response properties of lateral intraparietal area neurons at the single trial, single cell level; the capability of GLM to capture a broad range of single neuron response behaviors was analyzed in6. In all these works, however, the focus was on the response of neurons to stimuli of different spatio- temporal complexity; even where network interactions were accounted for, they proved to be, for the overall dynamics displayed by the ensemble, an important, yet not decisive 10 FIG. 5: Spatial microscopic network dynamics, data vs model. A Spatial distribution of the average time-rank of neuron activations during bursts for data and model. Averages are computed over all bursts in the data and a simulation of comparable length. B Spatial distribution of time-rank for 2 single bursts in data. C Spatial distribution of time-rank for the 2 bursts from simulation, most similar to those showed in B. D Distribution of the difference of the similarity between bursts from data, and both bursts from simulation and bursts from surrogate simulated data. E Scatterplot of post-synaptic efficacies of the neuron with smallest time-rank vs the rank of its post-synaptic targets. correction. To our knowledge, no published study to date has focused on the autonomous dynamics of GLM networks inferred from neuronal data. Important progress has also been made on the ability of GLMs to predict single neuron spiking in an ensemble of neurons, with accuracy higher than that provided by methods based on the instantaneous state of the ensemble itself2 and the PSTH4. In the lingo of the present paper, though, these attempts were based on simulations of the inferred network in driven mode, with at most the activity of a single neuron free to reenter just that single 11 neuron's dynamics. To our knowledge, our work for the first time has attempted a direct, microscopic comparison between the activity autonomously generated by an inferred GLM network and the one observed in the biological data. The free mode of simulation opens the way, in principle, to multi-time-step, whole network activity predictions. Being the model probabilistic and the data noisy, it is of course expected that the ability of the model, in free simulation mode after having been driven by the data for some time, to follow the experimentally observed spike trains should quickly deteriorate. In fact, this is what we found with our model: a burst becomes practically unforeseeable going some 100-200 ms in the past. This result is in contrast with the impressive findings reported in29, in which the authors, on recordings similar to the ones used in this work, through a new model- free method based on state-space reconstruction, were able to predict the occurrence of a burst even 1 s in advance. Whilst it is not clear why this is the case, we applied the state-reconstruction method on our data with negative results, in agreement with what we found with our model. We presented several semi-quantitative comparisons between the spike trains generated by the model network, without any time-varying external input, and the data from which the model's parameters have been inferred. Such tests, albeit successful, represent of course just an example of the many possible ways to compare two bursting dynamics. Yet we stress how a similar model, that has been widely and successfully applied to the analysis of neuronal data and has a slightly greater number of parameters too, failed in large part to reproduce the most prominent features of the dynamics displayed by the data. Two interrelated questions pertain to how representative the chosen experimental system is of neuronal network dynamics in general and the capability of the presented framework to generalize to other kinds of non-stationary neural activities. As for the first question, our primary concern in selecting an ex-vivo bursting culture has been to study a system that, absent any external stimuli, would show a non-trivial collective behavior. As dis- cussed above, the capability of GLMs to mimic the response properties of single neurons or a network to stimuli has already been well investigated. On the other hand, numerous studies in literature witness to the complexity of spatial and temporal behavior of neu- ronal cultures8–20. As for the second question, we remark that our primary goal was to devise a 'proof of concept' extension to show that a GLM can autonomously generate a complex non-stationary activity, and not to establish general conditions for such capability. We therefore chose a bursting regime as a testbed for the model not because it is exem- plary, but because it represented a clear challenge: a collective, non-linear phenomenon that demands the model to adapt to a broad spectrum of time- and space-scales. There were of course sensible alternatives to choose from, the most relevant being, e.g., the UP and DOWN dynamics in brain slices; although we have not tackled the issue directly, our results (see Fig. 3B, bottom plot) hint at a inherent capability of the model to sustain a noisy bistable dynamics, as more in general suggested in30. Nevertheless, the extent to which generative models are able to adapt to different regimes, and how dynamic network effects and stimulation characteristics interact in a generative model like the one studied here are open issues, that will probably attract much interest in the near future. Given our goal to establish the capability of a GLM to reproduce a complex, non-linear dynamics, we found two ingredients - a non-symmetric sigmoidal transfer function and a Negative Binomial spike-generation model - that proved to be minimal: absent one of them, the model's performances were drastically impaired. Even if we do not claim these specific ingredients to be necessary in general, we are persuaded they represent instances of mechanisms that are both biologically plausible and computationally desirable for the model. Although a sigmoidal, saturating transfer function would appear to be a natural choice for a model meant to reproduce neural activities, surprisingly this option has never been explored, to our knowledge, to perform network inference on neuronal recordings. While the use of an exponential transfer function in GLMs is grounded in general statistical requirements31, in our case an asymmetric sigmoid appears to be the single most important factor explaining the success of the proposed model. As already noted, the saturation of the sigmoidal function allows naturally for a bimodal distribution of firing rates and thus makes 12 model's behavior more robust in the face of the intrinsic fluctuations of network activity. Therefore, if the specific form we chose is just one possible instance among many plausible ones, a saturating behavior is expected to benefit the model beyond the immediate scope of the present work; a certain degree of asymmetry also appears to be beneficial according to our results, allowing for differential sensitivity to high and low input levels. It is interesting to note how this finding contrast with the success of non-saturating transfer functions in deep learning literature, where the introduction of rectified linear units, in place of the standard logistic ones, has represented one of the major breakthroughs of recent years32. Such units exhibit 'intensity equivariance', that is the ability to readily generalize to data points that differ only for a scale factor; whilst such property is clearly valuable when dealing with data such natural images and sounds, when applying machine learning techniques to very noise and sparse data, such as in the case studied here, bounded transfer functions are probably beneficial exactly for the opposite reason: they filter out most of the incoming information to gain a poorer but more robust coding in the output. We have seen that, if the role of a sigmoid gain function seems to facilitate a bistable behavior of the network, the negative binomial's one is to efficiently destabilize, with large fluctuations of the spike counts, the high-activity state. Recent experimental evidence has emerged supporting the Negative Binomial distribution as a candidate for the spike counts variability in real neurons25; moreover, a negative binomial spike generation has already been adopted in a model combining GLM with flexible graph-theoretic priors for the connectivity33. Although, then, our second ingredient already finds experimental and theoretical support in the literature, we provide a new hypothesis on why the super-Poisson statistics arises: as the effect of input fluctuations generated by the activity of neurons that have not been recorded. And our hypothesis directly hints to the Negative Binomial as just one possible way to model such effect, where over-dispersed spike counts are, instead, a necessary signature of it; it is this more general feature, in our opinion, that proved to be so important for the success of the proposed model. Interestingly, our findings resonate with the recently proposed role of fluctuating unobserved variables in the emergence of criticality in a wide range of systems34. Our hypothesis presupposes that the recorded neurons have been chosen at random from the whole network; this is probably not the case in many instances, where a local set of units is sampled instead; such non-random sub-sampling can potentially lead to strong deviations from expected behavior23; this of course could produce systematic effects in the inference. Besides, our model cannot incorporate the effects of slow changes in single neuron excitability, as observed experimentally in35,36, and analyzed in a statistical model in25. METHODS Experimental data As originally described in10, cortical neurons were obtained from newborn rats within 24 hours after birth, following standard procedures. Briefly, the neurons were plated directly onto a substrate-integrated multielectrode array (MEA). The cells were bathed in MEM supplemented with heat-inactivated horse serum (5%), glutamine (0.5 mM), glucose (20 mM), and gentamycin (10 µg/ml) and were maintained in an atmosphere of 37◦ C, 5% CO2/95% air in a tissue culture incubator as well as during the recording phases. The data analyzed here was collected during the third week after plating, thus allowing functional and structural maturation of the neurons. MEAs of 60 Ti/Au/TiN electrodes, 30 µm in diameter, and spaced 200 µm from each other (Multi Channel Systems, Reutlingen, Ger- many) were used. The insulation layer (silicon nitride) was pretreated with poly-D-lysine. All experiments were conducted under a slow perfusion system with perfusion rates of ∼100 µl/h. A commercial 60-channel amplifier (B-MEA-1060; Multi Channel Systems) with fre- quency limits of 1-5000 Hz and a gain of 1024× was used. The B-MEA-1060 was connected to MCPPlus variable gain filter amplifiers (Alpha Omega, Nazareth, Israel) for additional 13 amplification. Data was digitized using two parallel 5200a/526 analog-to-digital boards (Mi- crostar Laboratories, Bellevue, WA). Each channel was sampled at a frequency of 24000 Hz and prepared for analysis using the AlphaMap interface (Alpha Omega). Thresholds (8× root mean square units; typically in the range of 10-20 µV ) were defined separately for each of the recording channels before the beginning of the experiment. The electrophysiological data are freely available from S. Marom (http://marom.net.technion.ac.il/). ACKNOWLEDGEMENTS We are grateful to Shimon Marom for having shared his data, and for several illuminating discussions along the way. 1W. Truccolo, U. T. Eden, M. R. Fellows, J. P. Donoghue, and E. N. Brown, "A point process framework for relating neural spiking activity to spiking history, neural ensemble, and extrinsic covariate effects," Journal of neurophysiology 93, 1074–1089 (2005). 2W. Truccolo, L. R. Hochberg, and J. P. Donoghue, "Collective dynamics in human and monkey sensori- motor cortex: predicting single neuron spikes," Nature Neuroscience 13, 105–111 (2010). 3C. Capone, C. Filosa, G. Gigante, F. Ricci-Tersenghi, and P. Del Giudice, "Inferring synaptic structure in presence of neural interaction time scales," PloS one 10, e0118412 (2015). 4J. W. Pillow, J. Shlens, L. Paninski, A. Sher, A. M. Litke, E. Chichilnisky, and E. P. Simoncelli, "Spatio- temporal correlations and visual signalling in a complete neuronal population," Nature 454, 995–999 (2008). 5I. H. Stevenson, J. M. Rebesco, N. G. Hatsopoulos, Z. Haga, L. E. Miller, and K. P. Kording, "Bayesian inference of functional connectivity and network structure from spikes," IEEE Transactions on Neural Systems and Rehabilitation Engineering 17, 203–213 (2009). 6A. I. Weber and J. W. Pillow, "Capturing the dynamical repertoire of single neurons with generalized linear models," Neural Computation 29, 3260–3289 (2017). 7I. M. Park, M. L. Meister, A. C. Huk, and J. W. Pillow, "Encoding and decoding in parietal cortex during sensorimotor decision-making," Nature neuroscience 17, 1395–1403 (2014). 8G. Gigante, G. Deco, S. Marom, and P. Del Giudice, "Network events on multiple space and time scales in cultured neural networks and in a stochastic rate model," PLoS computational biology 11, e1004547 (2015). 9T. Baltz, A. Herzog, and T. Voigt, "Slow oscillating population activity in developing cortical networks: models and experimental results." J Neurophysiol 106, 1500–1514 (2011). 10D. Eytan and S. Marom, "Dynamics and effective topology underlying synchronization in networks of cortical neurons," The Journal of Neuroscience 26, 8465–8476 (2006). 11M. Giugliano, P. Darbon, M. Arsiero, H.-R. Luscher, and J. Streit, "Single-neuron discharge properties and network activity in dissociated cultures of neocortex," Journal of Neurophysiology 92, 977–996 (2004). 12T. Gritsun, J. le Feber, J. Stegenga, and W. L. Rutten, "Experimental analysis and computational mod- eling of interburst intervals in spontaneous activity of cortical neuronal culture," Biological Cybernetics 105, 197–210 (2011). 13I. P. I. Park, D. X. D. Xu, T. B. DeMarse, and J. C. Principe, "Modeling of Synchronized Burst in Dissociated Cortical Tissue: An Exploration of Parameter Space," (2006). 14D. A. Wagenaar, J. Pine, and S. M. Potter, "An extremely rich repertoire of bursting patterns during the development of cortical cultures," BMC neuroscience 7, 11 (2006). 15J. M. Beggs and D. Plenz, "Neuronal avalanches in neocortical circuits," The Journal of Neuroscience 23, 11167–11177 (2003). 16T. Petermann, T. C. Thiagarajan, M. A. Lebedev, M. A. Nicolelis, D. R. Chialvo, and D. Plenz, "Sponta- neous cortical activity in awake monkeys composed of neuronal avalanches," Proceedings of the National Academy of Sciences 106, 15921–15926 (2009). 17D. Plenz and T. C. Thiagarajan, "The organizing principles of neuronal avalanches: cell assemblies in the cortex?" Trends in neurosciences 30, 101–110 (2007). 18D. Plenz and H. G. Schuster, Criticality in neural systems (Wiley-VCH New York, NY, 2014). 19F. Lombardi, H. J. Herrmann, D. Plenz, and L. de Arcangelis, "Temporal correlations in neuronal avalanche occurrence," Scientific reports 6, 24690 (2016). 20M. Yaghoubi, T. de Graaf, J. G. Orlandi, F. Girotto, M. A. Colicos, and J. Davidsen, "Neuronal avalanche dynamics indicates different universality classes in neuronal cultures," Scientific reports 8, 3417 (2018). 21T. Masquelier and G. Deco, "Network bursting dynamics in excitatory cortical neuron cultures results from the combination of different adaptive mechanism," PloS one 8, e75824 (2013). 22A. Peyrache, N. Dehghani, E. N. Eskandar, J. R. Madsen, W. S. Anderson, J. A. Donoghue, L. R. Hochberg, E. Halgren, S. S. Cash, and A. Destexhe, "Spatiotemporal dynamics of neocortical excitation and inhibition during human sleep," Proceedings of the National Academy of Sciences 109, 1731–1736 (2012). 23A. Levina and V. Priesemann, "Subsampling scaling," Nature communications 8, 15140 (2017). 14 24A. Onken, S. Grunewalder, M. H. Munk, and K. Obermayer, "Analyzing short-term noise dependencies of spike-counts in macaque prefrontal cortex using copulas and the flashlight transformation," PLoS computational biology 5, e1000577 (2009). 25R. L. Goris, J. A. Movshon, and E. P. Simoncelli, "Partitioning neuronal variability," Nature neuroscience 17, 858–865 (2014). 26Y. Roudi, B. Dunn, and J. Hertz, "Multi-neuronal activity and functional connectivity in cell assemblies," Current opinion in neurobiology 32, 38–44 (2015). 27J. Tyrcha, Y. Roudi, M. Marsili, and J. Hertz, "The effect of nonstationarity on models inferred from neural data," Journal of Statistical Mechanics: Theory and Experiment 2013, P03005 (2013). 28O. M. A. D. U. F. Trang-Anh Nghiem, Bartosz Telenczuk, "Maximum entropy models reveal the corre- lation structure in cortical neural activity during wakefulness and sleep," ArXiv (2018). 29S. Tajima, T. Mita, D. J. Bakkum, H. Takahashi, and T. Toyoizumi, "Locally embedded presages of global network bursts," arXiv preprint arXiv:1703.04176 (2017). 30V. Rostami, P. P. Mana, S. Grun, and M. Helias, "Bistability, non-ergodicity, and inhibition in pairwise maximum-entropy models," PLOS Computational Biology 13, e1005762 (2017). 31J. K. Lindsey, Applying generalized linear models (Springer Science & Business Media, 2000). 32V. Nair and G. E. Hinton, "Rectified linear units improve restricted boltzmann machines," in Proceedings of the 27th international conference on machine learning (ICML-10) (2010) pp. 807–814. 33S. Linderman, R. P. Adams, and J. W. Pillow, "Bayesian latent structure discovery from multi-neuron recordings," in Advances in Neural Information Processing Systems (2016) pp. 2002–2010. 34D. J. Schwab, I. Nemenman, and P. Mehta, "Zipfs law and criticality in multivariate data without fine-tuning," Physical review letters 113, 068102 (2014). 35A. Gal, D. Eytan, A. Wallach, M. Sandler, J. Schiller, and S. Marom, "Dynamics of excitability over extended timescales in cultured cortical neurons," The Journal of Neuroscience 30, 16332–16342 (2010). 36A. Gal and S. Marom, "Self-organized criticality in single-neuron excitability," Physical Review E 88, 062717 (2013).
1604.01680
4
1604
2016-11-09T09:50:39
The Complex Hierarchical Topology of EEG Functional Connectivity
[ "q-bio.NC" ]
Understanding the complex hierarchical topology of functional brain networks is a key aspect of functional connectivity research. Such topics are obscured by the widespread use of sparse binary network models which are fundamentally different to the complete weighted networks derived from functional connectivity. We introduce two techniques to probe the hierarchical complexity of topologies. Firstly, a new metric to measure hierarchical complexity; secondly, a Weighted Complex Hierarchy (WCH) model. To thoroughly evaluate our techniques, we generalise sparse binary network archetypes to weighted forms and explore the main topological features of brain networks- integration, regularity and modularity- using curves over density. By controlling the parameters of our model, the highest complexity is found to arise between a random topology and a strict 'class-based' topology. Further, the model has equivalent complexity to EEG phase-lag networks at peak performance. Hierarchical complexity attains greater magnitude and range of differences between different networks than the previous commonly used complexity metric and our WCH model offers a much broader range of network topology than the standard scale-free and small-world models at a full range of densities. Our metric and model provide a rigorous characterisation of hierarchical complexity. Importantly, our framework shows a scale of complexity arising between 'all nodes are equal' topologies at one extreme and 'strict class-based' topologies at the other.
q-bio.NC
q-bio
The Complex Hierarchical Topology of EEG Functional Connectivity Keith Smith∗,a,b, Javier Escuderoa aInstitute for Digital Communications, School of Engineering, University of Edinburgh, West Mains Rd, Edinburgh, EH9 3FB, UK, bAlzheimer Scotland Dementia Research Centre, University of Edinburgh, 7 George Square, Edinburgh, EH8 9JZ, UK e-mail: [email protected], [email protected] 6 1 0 2 v o N 9 ] . C N o i b - q [ 4 v 0 8 6 1 0 . 4 0 6 1 : v i X r a Abstract Background : Understanding the complex hierarchical topology of functional brain networks is a key aspect of func- tional connectivity research. Such topics are obscured by the widespread use of sparse binary network models which are fundamentally different to the complete weighted networks derived from functional connectivity. New Methods: We introduce two techniques to probe the hierarchical complexity of topologies. Firstly, a new metric to measure hierarchical complexity; secondly, a Weighted Complex Hierarchy (WCH) model. To thoroughly evaluate our techniques, we generalise sparse binary network archetypes to weighted forms and explore the main topological features of brain networks- integration, regularity and modularity- using curves over density. Results: By controlling the parameters of our model, the highest complexity is found to arise between a random topology and a strict 'class-based' topology. Further, the model has equivalent complexity to EEG phase-lag networks at peak performance. Comparison to existing methods: Hierarchical complexity attains greater magnitude and range of differences between different networks than the previous commonly used complexity metric and our WCH model offers a much broader range of network topology than the standard scale-free and small-world models at a full range of densities. Conclusions: Our metric and model provide a rigorous characterisation of hierarchical complexity. Importantly, our framework shows a scale of complexity arising between 'all nodes are equal' topologies at one extreme and 'strict class-based' topologies at the other. Keywords: Functional connectivity, Hierarchical complexity, Brain networks, Electroencephalogram, Network simulation 1. Introduction Graph theory is an important tool in functional con- nectivity research for understanding the interdependent activity occurring over multivariate brain signals [1–3]. In this setting, Complete Weighted Networks (CWNs) are produced from all common recording platforms including the Electroencephalogram (EEG), the Magnetoencephalo- gram (MEG) and functional Magnetic Resonance Imaging (fMRI), where every pair of nodes in the network share a connection whose weight is the output of some connectiv- ity measure. Complex hierarchical structures are known to exist in real networks [4], including brain networks [1, 5], for this reason it is important to find methods to specifi- cally evaluate hierarchical complexity of network topology. Here we introduce methods specific to this end. Complexity is understood neither to mean regularity, where obvious patterns and repetition are evident, nor randomness, where no pattern or repetition can be estab- lished, but attributed to systems in which patterns are irregular and unpredictable such as in many real world phenomena [6]. Particularly, the brain is noted to be such a complex system [7] and this is partly attributed to its hierarchical structure [5]. Hierarchical complexity is thus concerned with understanding how the hierarchy of the system contributes to its complexity. Here we introduce a new metric aptly named hierarchical complexity, R, which is based on targeting the structural consistency at each hierarchical level of network topology. We compare our metric with network entropy [8] and find that we can of- fer a greater magnitude and density range for establishing differences in complexity of different graph topologies. Alongside this, we introduce the Weighted Complex Hi- erarchy (WCH) model which simulates hierarchical struc- tures of weighted networks. This model works by mod- ifying uniform random weights by addition of multiples of a constant, which is essentially a weighted preferential selection method with a highly unpredictable component provided by the original random weights. We show that it follows very similar topological characteristics of networks formed from EEG phase-lag connectivity. Intrinsic to our model is a strict control of weight ranges for hierarchi- cal levels which offers unprecedented ease, flexibility and rigour for topological comparisons in applied settings and for simulations in technical exploration for brain network analysis. This also provides an unconvoluted alternative to methods which randomise connections [9, 10] or weights Preprint submitted to Journal of Neuroscience Methods November 10, 2016 [11] of the original network. Any rigorous evaluation of brain networks should ad- dress their inherent complete weighted formulation [12]. However, the current field has largely lacked any concerted effort to build an analytical framework specifically tar- geted at CWNs, preferring instead to manipulate the func- tional connectivity CWNs into sparse binary form (e.g. [10, 13, 14] as well as wide-spread use of the Watts-Strogatz [9] and Albert-Barabasi [15] models) and using the pre- existing framework built around other research areas which have different aims and strategies in mind [16]. In our methodological approach we propose novel generalisations of pre-existing sparse binary models to CWN form and thus allow a full density range comparison of our tech- niques. Due to the intrinsic properties of these graph types we find minimal and maximal topologies which can help to shed light on a wide variety of topological forms and their possible limitations [8] in a dense weighted framework. Further, as part of our study we seek after straightfor- ward metrics to evaluate other main aspects of network topology for comparisons [8, 17] and, in this search, found it necessary to revise key network concepts of integration- segregation [2, 9, 18] and scale-freeness [15, 19]. We pro- vide here these revisions: i) That the clustering coefficient, C, is enough to analyse the scale of integration and seg- regation, finding it unnecessary and convoluted to use the characteristic path length, L, as a measure of its opposite, as generally accepted [1, 9]. ii) We provide mathematical justification that the degree variance, V , and thus network irregularity [20] is a strong indicator of the scale-free factor of a topology. Our study of hierarchical complexity, using a compre- hensive methodological approach, provides mathematical quantification of the hierarchical complexity of EEG func- tional connectivity networks and reveals new insights into key aspects of network topology in general. Our model provides improved comparative abilities for future clinical and technical research. 2. Network Science: Proposed methods and key revisions We adopt the notation in [21] so that a graph, G(V, W), is a set of n nodes, V, connected according to an n × n weighted adjacency matrix, W. Entry Wij of W corre- sponds to the weight of the connection from node i to node j and can be zero. An unweighted graph is one in which connections are distinguished only by their existence or non-existence, so that, without loss of generality, all ex- isting connections have weight 1 and non-existent connec- tions have weight 0. The graph is undirected if connec- tions are symmetric, which gives symmetric W. A simple graph is unweighted, undirected, with no connections from a node to itself and with no more than one connection be- tween any pair of nodes. This corresponds to a graph with a symmetric binary adjacency matrix with zero di- agonal. Such graphs are easy to study and measure [16]. ki =(cid:80)n The degree, ki, of node i is defined as the number of its adjacent connections, which is the number of non zero en- tries of the ith column of W. Then, for a simple graph, j=1 Wij. For a graph with 2m edges, the connec- tion density, P , of a graph is P = 2m/n(n − 1). A CWN is represented by a symmetric adjacency ma- trix with zero diagonal (no self-loops) and weights, Wij ∈ [0, 1], elsewhere. To analyse CWNs it is beneficial to con- vert it to simple form by binarising the adjacency matrix using a threshold, where a percentage of strongest con- nections are set to 1 and the remaining values set to 0. This stays true to the network activity [12] whilst reducing computational complexity and weight issues found with weighted metrics [2]. Hereafter, all mathematics will refer to simple graphs. In this section we present the contributions of this study. We first present the hierarchical complexity metric and the WCH model, which are the key novel contributions of this paper. Thereafter we detail revisions and clarifica- tion of integration and segregation as a scale evaluated by C and scale-freeness as a factor evaluated by V . Finally, we outline the generalisation of key network archetypes to CWN form, full details of which can be found in the supplementary material. 2.1. Hierarchical Complexity Metric The ideas of order and complexity are well known in the discussion of networks (indeed, real world networks are often called complex networks [1, 3, 22]). In math- ematics, the graphs studied derive from some theoretical principles. These can involve set patterns, without ran- dom fluctuations of connections, such as regular networks, fractal networks, star networks and grid networks. On the other hand much interest is shown in more randomly gen- erated topologies, such as random graphs and other graphs involving random processes, as these express something of the more erratic and irregular quality of connections in networks constructed from real world phenomena [9, 23]. However, real world phenomena differ from random pro- cesses in that there is a clear organisational behaviour ap- parent throughout the hierarchical structure, both within hierarchical levels and between hierarchical levels [4, 5]. Although this structure is perhaps impossible to retrace, because its formation inevitably involves many unknown generative processes, we can provide methods for its anal- ysis. Hierarchies in networks are generally determined by degrees of nodes, where a small group of highly connected nodes create a rich club [22] on the top hierarchical level and nodes with generally lesser connectivity exist on a pe- ripheral lower levels. Further, it is seen that a node's re- lationship within the context of the network is greatly de- termined by the other nodes to which it is connected [24]. Thus, to understand the hierarchical complexity of a net- work we propose to study the behaviour of nodes of a given degree by looking at the degrees of nodes in their neigh- bourhoods. We define D as the set of degrees of a graph, 2 disorganised or more complexly organised. For example, in Fig.1.B the two degree nodes all have the same degree sequences- {3, 4}- whereas the three degree nodes are split into two different degree sequences- {1, 2, 2} and {1, 1, 4}- and finally the neighbourhood degree sequences of the four degree nodes are all different- {1, 1, 1, 4}, {1, 2, 2, 4} and {2, 3, 3, 3}. So the complexity of just the two degree nodes is 0, the complexity of just the three degree nodes is 2((2− 1.5)2 + (1 − 1.5)2 + (2 − 3)2 + (4 − 3)2)/(4 × 3 × 3) = 5/36 and the complexity of just the four degree nodes is (2(1 − 4/3)2 + (2 − 4/3)2 + 2((1 − 2)2 + (3 − 2)2) + 2(4 − 11/3)2 + (3 − 11/3)2)/(4 × 3 × 2) = (16/3)/24 = 8/36, the complexity over all three levels being the average- 13/108. This measure is thus minimal for graphs in which, for each k and k(cid:48), every k-degree node is connected to exactly the same number of k(cid:48)-degree nodes. This property, for ex- ample, is seen in ring lattices, and quasi-star graphs and is close to minimal in the line graph, fractal graphs and grid lattices. Furthermore, the degrees of random networks are known to have a fairly small spread which is a factor pe- nalised by our complexity value. Thus random networks should obtain low values of our complexity measure. On the other hand, R values of real networks are expected to be higher given the high spread and degree fluctuations of those networks caused by hub nodes promoting a high degree irregularity while the spontaneous nature of real- world connections should promote a high variability of the neighbourhood degree sequences. 2.2. Weighted Complex Hierarchy Model The foundation of our model is the random CWN model. The most general random network is the Erdos-R´enyi (E- R) random network [23] which is formed by assigning a probability, p, to the question of the existence or non- existence of connections on a network with n nodes. Such a construct is, in fact, an ensemble of graphs denoted G(n, p). A sample of this ensemble is obtained by generat- ing a random value for every possible connection and ap- plying the probability value p as a threshold to see whether or not that connection should exist in our sample. The ran- dom CWN model is thus simply a symmetric matrix with zero diagonal and randomly generated values Wij ∈ [0, 1] elsewhere. If we threshold the CWN at weight T = p, we recover a binary Erdos-R´enyi random graph from the random graph ensemble G(n, p). Starting from an Erdos-R´enyi CWN we randomly dis- tribute the nodes into hierarchy levels based on some dis- crete cumulative distribution function, p, by generating a random number, r, between 0 and 1 for each node and putting the node in the level for which r − p is first less than 0. We then distribute ls additional weight to all connections of adjacent nodes in the lth level, for some suitably chosen s. The parameters of this model are then (n, s, l, p). The parameter n is the number of nodes in the network. The parameter s is the strength parameter, which is constant since the random generation of the ini- tial weights is enough to contribute to weight randomness. Figure 1: A. Example of a node degree neighbourhood. Here is shown a part of a network relating to the neighbourhood of the blue node. The blue node has neighbourhood degree sequence {1, 2, 3, 4, 4}, i.e. the ordered degrees of the orange nodes. Grey connections indicate all the additional connections of the orange nodes in the network. B. Example for graph complexity. Here is shown a 20 node network with varying 'orderedness' at different degree levels. C. Diagram of the construction of the WCH model. Above is the probability distribution function for a geometric distribution with p = 0.6 for a three level hierarchy. Below is a graphic displaying the additional weight added between nodes in given hierarchy levels. G. Similar to the idea of node degree sequences [25], we can construct neighbourhood degree sequences, specific to each node in the graph. That is, for a node, i, of degree k ∈ D we have a sequence si = {di,1, di,2, , di,k} s.t. di,1 ≤ di,2 ≤ ··· ≤ di,k ∈ D, where di,j is the degree of the jth node connected to node i (see Fig. 1.A). For all nodes of a given degree, k, the corresponding neighbourhood degree sequences have equal length, k. We define the hierarchical complexity, R, of a network as the average variance of the k-degree neighbourhood de- gree sequences and can be expressed as: (cid:88) Dk(cid:54)=∅ R = 1 D 1 krk(rk − 1)  k(cid:88) j=1 (cid:32)(cid:88) i∈Dk (cid:33) , (ski(j) − µkj)2 (1) where D is the number of distinct degrees in the graph, Dk is the set of nodes of degree k, ski(j) is the jth element of the ith k-length sequence, µkj is the mean value of el- ement j over all k-length sequences and rk is the number of nodes of degree k, which is added to the denominator for normalisation of hierarchy levels. Organisation of the graph at the level of k-degree nodes can be seen by comparing the jth elements of their neigh- bourhood sequences. If all of the jth elements of all the sequences are equal, that is si = sj for all si, sj of length k, then there is a high degree of order present in the k-degree nodes of the graph. If these sequences differ widely how- ever, then it can be said that the k-degree nodes are either 3 The parameter l is the number of levels of the hierarchy, with a default setting of a random integer between 2 and 5. The vector p is the cumulative probability distribution vector denoting the probabilities that a given node will be- long to a given level where the default, which we use here, is a geometric distribution with p = 0.6 in hierarchical levels (0, 1, 2, . . . , l) where the nodes with highest connec- tivity (top hierarchical level) are at the tail end of the distribution. Fig. 1.C plots an example of the geometric distribution for a three level hierarchy. The text inside the box plots, above, indicates the additional weights given to connections adjacent to nodes inside the given level. The graphic below explains the additional weights provided by the strength parameter of connections between nodes in different levels as well as in the same level. For example, a connection between a level 1 node and a level 2 node has additional strength 3s which consists of one s provided by the node in Level 1 and 2s provided by the node in Level 2. At s = 0, we have the E-R random network and at s = 1 the weights of the network are linearly separable by the hi- erarchical structure producing a strict 'class-based' topol- ogy. Between these values a spontaneous 'class-influenced' topology emerges. 2.3. Revision of concepts from network science Here we present justifications for metrics as measures of key topological factors- the global clustering coefficient, C, for degree of segregation and the degree variance, V , for irregularity, linked to scale-freeness. 2.3.1. Integration-Segregation The concept of integration in brain networks is closely tied in to the small world phenomenon [26], where real world networks are found to have an efficient 'trade off' between segregative and integrative behaviours [27]. The most widely used topological metrics in network science- C and the characteristic path length, L- are commonly noted as measures of these quantities, respectively. Here, L is defined as the average of the shortest paths between each pair of nodes and C is defined as the probability that a path of length 2, or triple, in the graph has a shortest path of length 1. That is, C = closed triples triples , (2) where a closed triples is such that, for triple {Wik, Wkj}, Wij = 1, for i, j, k distinct. Since integration implies a non-discriminative behaviour in choice, we argue that the random graph ensemble [23], defined by its equal probability of existent connections be- tween all pairs of nodes, is the most exemplary model of an integrated network. Anything which deviates from equal probability is a discriminative factor which favours certain connections or nodes over others, likely leading to more segregated activity. Further, it is clear that integration and segregation are opposite ends of the same spectrum- 4 something which is not integrated must be segregated and vice versa. Having one metric to inform on where a net- work lies on that spectrum is therefore sufficient. Thus, here we propose C as the topological measure to evaluate levels of integration (and so segregation) of a given network. Firstly, we note that values of C for random graphs and small-world graphs are often much more dis- tinguishable than those of L [9] and it is certainly assumed that these graphs have very different levels of integration. Secondly, since the random network is optimally integrated and E[Cran] = E[Pran] [16], where Pran is the connection density of the random network, then the larger the devi- ation from 1 of the value γ = C/E[Cran] = C/E[Pran] = C/P , the more segregated is the network. We will include both L and C in our analysis in order to provide evidence to back the above proposal. 2.3.2. Regularity and Scale-Freeness Another topological factor of small world networks is noted as a scale-free nature characterised by a power law degree distribution [28]. To understand this aspect of net- work topology another factor of network behaviour is for- mulated distinguishing between 'line' like and 'star' like graphs [2, 29]. Here, we show that characterisation of scale-freeness is closely connected to the regularity of a network. Regular graphs have been studied for over a century [30]. They are defined as graphs for which every node has the same degree. An almost regular graph is a graph for which the highest and lowest degree differs by only 1. Thus a highly irregular graph can be thought of as any graph whose ver- tices have a high variability. Such behaviour can be cap- tured simply by the variance of the degrees present in the graph, that is (3) where D = {ki}i∈V , is the set of node degrees on a given graph [20]. V = var(D), graph with degrees k = {k1, k2, . . . , kn}, and (cid:80)n For regular graphs V = 0 by definition, but more prob- ing is necessary to distinguish high V topology. For a i=1 ki = 2m, on multiplying out the brackets V simplifies to (cid:19)2 − ki (cid:18) 2m n(cid:88) i=1 n − 2mP, V = = 1 (n − 1) (cid:107)k(cid:107)2 (n − 1) 2 2 = (cid:80)n i=1 k2 where P = 2m/n(n − 1) is the connection density and (cid:107)k(cid:107)2 i , is the squared (cid:96)2 norm of k. This tells us that V is proportional to the sum of the squares of the degrees of the graph, (cid:107)k(cid:107)2 2, and, for fixed number of connections, m, V in fact depends only on (cid:107)k(cid:107)2 2. Now, it is known that (cid:107)k(cid:107)2 2 is maximal in quasi-star graphs and quasi-complete graphs [31]. Essentially, the quasi-star graph has a maximal number of maximum degree nodes in the graph for the given connection density and the quasi- complete graph has a maximal number of isolated, or zero- degree, nodes in the graph. This tells us that, for low P , high V denotes the presence of a few high degree nodes and a majority of relatively low degree nodes, i.e. scale- free-like graphs. Thus, due to the restriction placed on possible degree distributions by the number of edges (the small number of edges in sparse networks means the num- ber of high degree nodes is very limited), the irregularity of degrees is a strong indicator of the strength of 'decay' of the given distribution, relating to how 'scale-free' the graph is. 2.4. Complete Weighted Network Archetypes In the supplementary material we detail the method to generalise sparse binary network archetypes to CWN form. The pre-requisit of such a generalisation is that we require obvious higher density versions of lower density forms which can be arranged in adjacency matrix form such that each non-zero entry, Wij, of the lower density adjacency matrix exists as a non-zero entry, Wij in the higher density adjacency matrix. This is indeed the case for the Regular Ring Lattice, Star, Grid Lattice and Frac- tal Modular CWNs (see Fig. 2 A,B,C,D respectively). We explain these higher and lower density forms of the bi- narised CWN in terms of weight categories where, if we choose an appropriate threshold, T , we can recover all edges in the same and all higher weight categories and none of the edges existing in all lower categories. 3. Methods Here we apply methods to graphs of 64 nodes, typical of medium density EEG. For analysis we employ connec- tion density thresholds at integer percentages of strongest weighted connections, rounded to the nearest whole num- ber of connections. We then implement metric algorithms on each of these binary networks and plot the obtained values on a curve against connection density, similar as in e.g. [32, 33]. This generates metric curves plotted against connection density which provides a detailed analysis of the CWN topology. Other methods exist to analyse CWNs such as weighted metrics [11] or density integrated metrics [34], but these metrics still give only singular values for a given network which belies little of topological behaviour at different scales of connectivity strength. For random and WCH CWNs we use sample sizes of 100 for each network and for the EEG functional connec- tivity CWNs we have a sample size of 109 [35]. On the metric curves for these we plot the median with the in- terquartile range shaded in. For ordered networks there is only one network per type by definition. Our analytical framework is composed of a mixture of entirely new concepts and novel generalisations of ex- It is constituted of the isting concepts to CWN form. Figure 2: A. A 12 node ring lattice of degree 6, comprising the three strongest weight categories of the ring lattice CWN. B. The quasi-star with 4 nodes of degree n − 1 and n − 4 nodes of degree 4, also comprising the first four categories of the star CWN. C. The grid lattice weight categorisation (relating to the grey node) in a 30 node network (see supplementary material). Colours of edges denote category: black, blue, green, orange and red edges are in weight cate- gories 1, 2, 3, 4 & 5, respectively. The increasingly lighter boundaries thus represent 'catchment' areas around the node by increasing cat- egory. Centring these 'catchment' areas around a given node gives the respective categorisation of edges adjacent to the new node. D. Fractal modular CWN weight categorisation on 30 nodes. Edges shown (black) are 1st weight category edges. In this instance, in- creasingly lighter background represents areas within which all pairs of nodes become connected by edges when the network is subject to the threshold corresponding to the respectively increasing category (see supplementary material). following elements: four metrics, R, C, V , Q characteris- ing four important and distinct topological features; five CWN archetypal models- Random, Star, Regular Lattice, Fractal Modular, Grid Lattice; the WCH model. 3.1. Metrics In [8, 17] an 'architecture' of network topology is pro- posed involving the three most widely studied properties of brain networks- integration (and segregation) [2, 9, 18], 'scale-freeness' [15, 19] and modularity [5, 36]. For our analysis in comparison with hierarchical complexity, R, we choose a straightforward metric for each of these topo- logical factors- C for integration, V for scale-freeness and Q for modularity [36] where (cid:18) (cid:88) (cid:19) Q = 1 2m i,j Wij − kikj 2m δ(ci, cj), (4) where ci is the module containing node i and δ() is the Kronecker delta function. Highly efficient algorithms have been created [37, 38] aiming to maximise the value of Q for a given network. To compute the modularity of our networks, we use the undirected modularity function [37] in the Brain Connectivity Toolbox [18]. 5 3.2. Comparison for Hierarchical Complexity degree distribution qi = kipi/(cid:80) We compare our hierarchical complexity metric with a commonly used metric for analysing the entropy of the network degrees [8]. This is defined using the normalised j kjpj, where ki is the de- gree of node i and pi is the proportion of nodes in the graph with the same degree as node i which relates to probabilities of going to/ coming from neighbouring nodes in directed graph problems. Then the entropy of graph G is a straightforward derivation of Shannon's entropy equa- tion [39] for the degrees of the graph: H(G) = − n(cid:88) qi log(qi). (5) Thus, Network entropy encodes the eccentricity of the graph degrees. i=1 3.3. Comparisons for the WCH model We implement comparisons with the Watts-Strogatz small-world model [9] which randomly rewires a set pro- portion of edges starting from a regular lattice. We use the full range of parameters for initial degree specification (2 up to 62) and random rewiring parameters from 0.05 in steps of 0.05 up to 0.95. For each combination of parame- ters, 100 realisations of the model were computed and C, V , Q, and R were measured. We further compare with Albert-Barabasi's scale-free model [15] which begins with a graph consisting of core of highly connected nodes to which the rest of the nodes are added one by one with a set degree but paired by edges to randomly selected nodes. We use an initial number of nodes of 15 and the additional node's degree from 3 up to 14 in order to reach larger den- sities. 3.4. EEG networks We use an eyes open, resting EEG data set with 64 nodes. We report on networks created from the beta (12.5- 32Hz) band using coherence and the debiased Weighted Phase-Lag Index (dWPLI) in order to account for differ- ent possible types of EEG networks while reducing redun- dancy of similar topological forms found between the fre- quency bands (see supplementary material). The dataset, recorded using the BCI2000 instrumenta- tion system [40], was freely acquired from Physionet [35]. The signals were recorded from 64 electrodes placed in the main in accordance with the international 10-10 system. We took the eyes open resting state condition data, con- sisting of 1 minute of continuously streamed data which were partitioned into 1s epochs and averaged for each of 109 volunteers. FieldTrip [41] was used for pre-processing, frequency analysis and connectivity analysis to obtain the adjacency matrices of complete weighted networks. The 64 channels were re-referenced using an average reference, the multi- taper method was implemented from 0 seconds onwards 6 using Slepian sequences and 2Hz spectral smoothing. A 0.5Hz resolution was obtained using one second of zero padding. We chose to analyse the matrices obtained from both the coherence and the debiased Weighted Phase-Lag Index (dWPLI) [42] to look for differences between net- work topologies of zero and non-zero phase lag dependen- cies in the channels [43]. We treat the data of all tasks as a single dataset to allow for the variability of the EEG network topologies since we are not interested here in the tasks themselves but on the behaviour of general EEG net- works obtained from dWPLI and coherence. 3.5. Statistical analysis Due to the polynomial formulation of the complexity measure, producing a non-normal distribution, we com- pare metric distributions using the Wilcoxon rank sum test. The z-score is used to ascertain the magnitude and direction of the relationship of the distributions. 4. Results 4.1. Metric Comparisons Fig. 3 shows the metric curves (i.e. metric plotted against network density) for C, V , Q, R, L and H for all archetypes as well as for the EEG dWPLI (red shade) and coherence (blue shade) networks. From these plots we see experimental evidence of maximal and minimal topolo- gies for the given topological characteristics. These max- imal and minimal topologies are explained as the curves whose lines are consistently lowest or highest over all densi- ties. Fractal Modular networks (purple lines) are maximal for both C and Q (top left and centre left, respectively). This is to be expected since the modules are complete sub-networks with very few connections between modules, maximising Q. Further this minimises the number of open triples in the graph, maximising C, by restricting open triples to relating only to those few connections which do extend between modules. The star CWN (orange lines) acts as a maximal topology for V , as expected from the theory explained in Section 2, while being a minimal topol- ogy for L (bottom left). Regular graphs, such as the ring lattice network (blue lines), give 0 degree variance and hi- erarchical complexity, thus are minimal topologies of these features. The results of Fig. 3 for 30 node networks, found in the supplementary material, follow the same relation- ships, providing evidence that these features are indepen- dent of network size. Comparing the plots in Fig. 3 of C (top left) with L (bottom left) and R (centre right) with H (bottom right), it is immediately clear that L and H show extreme behaviour at low densities while remaining consistent at higher densities. This exemplifies how these metrics are aimed at analysis of sparse networks, where it appears that values can take a much greater range than for higher den- sity networks. To explore these comparisons further we perform sta- tistical analysis with Wilcoxon rank sum tests on the dif- ferences of distributions of metric values of EEG dWPLI and E-R random networks as well as of EEG dWPLI and EEG coherence networks (Fig 4). The results show that C (right) and R (left) attain a greater range over edge den- sity, P , of significant differences than their counterparts, L and H. Particularly, R distinguishes differences from 1% up to 44% densities in the EEG dWPLI and coherence comparison (solid blue line), whilst entropy only can dis- tinguish differences from 1% up to 27% (solid yellow line). Further, the z-scores indicate that in the range 1-27%, the differences found in R are greater than those found using H. Comparing the EEG dWPLI networks with E-R ran- dom networks (Fig 4, left, dashed lines), both metrics find differences at all levels, but the magnitude of difference found by R (blue) is consistently greater than those found by H (yellow). Thus, our metric outperforms entropy in both magnitude and range of differences found. Similarly, C finds a greater range and magnitude of dif- ferences than L, Fig 4 right. In fact, C discerns differences at all connection densities for the two comparisons, while L fails to find differences after 62% in comparing dWPLI and coherence networks (solid yellow line) and after 73% in comparing dWPLI and random networks dashed yellow line). Furthermore, L displays inverse differences at low densities (1-12%) compared to higher densities in the dW- PLI vs random comparison (dashed yellow line). This in- consistency is undesirable for translatability of integrative behaviour of network types from sparse networks to more dense networks. C does not suffer from such behaviour, displaying a constant relationship of metric values through the full range of densities (solid and dashed blue lines). Given these results, for the rest of our analysis, we will drop L and H and focus on the four proposed metric, C, V , Q and R. We must emphasise that this is taken purely in terms of the simplicity of explaining a general topological factor and does not mean that L and H are not useful for other purposes. 4.2. Weighted Complex Hierarchy Null Model Fig. 5 shows the mean results of C (top left), V (top right), Q (bottom left) and R (bottom right) over 100 realisations of each of the WCH models. We include a reduced number of strength parameters in the figure (s = 0.1, 0.2, . . . , 0.7) than those computed (s = 0.05, 0.1, . . . , 0.75) for greater clarity. Above 0.75 the parameter begins to sat- urate as the weights of the hierarchy levels tend to linear separability (linear separability occurs when s = 1 since 0s, 1s, 2s, ... then places the edge weights, originally in [0, 1], in disjoint ranges [0, 1], [1, 2], [2, 3], ...). We see that WCH networks (grey shaded lines) exhibit curve behaviour similar to the EEG networks and E-R random graphs (as in Fig. 3). The scale-free model (red error bars) also ex- hibits a similar behaviour, however in stark contrast, the small-world model (blue error bars) exhibits very different 7 Figure 3: Topological metric values for integration (C), regularity (V ), modularity (Q), hierarchical complexity (R), characteristic path length (L) and network entropy (H) against network density, P . Curves relate to network models as indicated in the legend (bottom right). Figure 4: Positive (negative) values indicate the contrasted distribu- tions exhibit the relationship provided in the legend (or its opposite). Zero indicates p-value insignificant at 5% level. a) The hierarchical complexity, R (blue), compared with network entropy, H (yellow). b) The clustering coefficient , C (blue), compared with characteristic path length, L (yellow). P is the network density. behaviours than those of the EEG or WCH networks, ex- hibiting a strong unsuitability for comparisons with EEG networks with much higher modularity and highly right skewed V curve (Fig. 5, top left) towards high densities as well as a similar right skew in R (bottom left) which is opposite to the left skew found for WCH and EEG net- work types. Although the scale-free model exhibits similar Figure 6: The z-statistics of distributions with significant differences from a Wilcoxon rank sum test. Positive (negative) values indicate the contrasted distributions exhibit the relationship provided in the legend (or its opposite). Zero indicates p-value insignificant at 5% level. P is the connection density. works do indeed exhibit greater complexity than the WCH model. The strong exception to this is an inability to dis- tinguish significant differences between the maximal com- plexity s = 0.3 WCH model and dWPLI networks within 7-23% densities (bold yellow line). Also, as the weight pa- rameter increases, the high plateaus previously mentioned begin to take effect as in the medium ranges of density the R values of the dWPLI networks and WCH model becomes more indistinguishable, with greater complexity found in the range 55-57% in the WCH model with s = 0.4 (green line). 4.3. Null model approaching EEG phase-lag networks Fig. 7 shows the values of the four topological features- complexity, integration, regularity and modularity for EEG dWPLI networks and the WCH network with strength pa- rameter 0.2. We see clearly that these networks behave very similarly with respect to the given metrics. The most obvious difference is that the modularity, Q, of dWPLI EEG networks is higher (bottom left). Also, as previously discussed, the dWPLI network complexity is greater than the WCH model, but it is still by far the most comparable model for complexity of those presented here. 5. Discussion 5.1. Complexity as revealed by weighted complex hierarchy model The behaviour demonstrated by the WCH model with respect to R indicates that high complexity arises from a hierarchical structure in which a greater degree of vari- ability is present in the rankings of weights with respect to hierarchy level. Too little difference between levels and the hierarchy is too weak to maintain complex interactions, too much difference between levels and the complexity of the hierarchy is dampened by a more ordered structure produced from the tendency towards linear separability of the edge weights enforced by the strength parameter. Thus, we provide evidence that topological complexity is Figure 5: The topological characterisation of network models by clus- tering coefficient, C, degree variance, V , modularity, Q and hierar- chical complexity, R, plotted against network density, P . Grey lines indicate mean values of the weighted complex hierarchy model with increasing light shade indicating increasing strength parameter from s = 0.1 in steps of 0.1 up to s = 0.7. Red errorbars indicate values of the Albert Barabasi scale-free model. Blue errorbars indicate values of the Watts Strogatz small-world model with increasingly light blue indicating increasing proportion of edges being randomly rewired. tendencies in topological metrics to the WCH and EEG networks, its range of values and densities is clearly very limited and so, therefore, its ability for topological refine- ment. By increasing the strength parameter of the WCH model we change the topology in a smooth fashion with decreas- ing integration, regularity and modularity (Fig. 5, top left, top right and bottom left, respectively). Interestingly, R (bottom left) rises with increasing strength parameter from s = 0.05 up to s = 0.3 where it takes its maximum values at densities ranging from 1-30% before falling again from s = 0.35 until s = 0.7. Further, above s = 0.3, the curves begin to deviate significantly from those of the EEG dWPLI networks, exhibiting greater plateaus of high complexity (lighter grey lines) which are more comparable with the EEG coherence networks. Interestingly, the complexity of the EEG dWPLI net- works appears to attain maximal values of R of all the networks studied here (Fig. 3). The only model which comes close is the WCH model (Fig. 7, bottom right). To clarify this observation we perform Wilcoxon rank sum tests on R values of the EEG dWPLI networks against that of the WCH model with strength parameters ranging from s = 0.2 up to s = 0.4, i.e. two steps before and after the maximal complexity setting of s = 0.3. The results are displayed in Fig. 6. In the vast majority of instances of strength parameter and density, the EEG dWPLI net- 8 are built independently from the brain networks [10, 15] or are constructed by the randomisation of connections of the networks being compared [9, 18], run into problems with density specification (in the case of independent models) and reproducibility (in both types of model). With the WCH model, we can simply create a bank of simulated CWNs which can be used throughout the study in exactly the same way as we use the functional connectivity CWNs. As an example of the power and elegance of the pro- posed model, say we want to find maximum spanning trees [45] of our brain networks and compare with a null model, then we simply take the maximum spanning trees of our null model. In contrast, in [14] they use a convoluted re- verse engineering process by assigning random weights to the connections of Watts-Strogatz small world networks (which are themselves of limited comparability to brain networks) and computing the MST from these resulting sparse weighted networks. Further, as seen in Fig. 1.C, for technical studies which rely on network simulations, the WCH model is built on parameters which can be altered to subtly change the re- sulting topology. This allows for sensitive analysis of a new techniques ability to distinguish subtle topological dif- ferences. Such paradigms are evident in clinical studies where, for example, one may try to distinguish between healthy and ill patients [29, 33] or between different cog- nitive tasks [46], so that this null model offers simulations which are directly relatable to clinical settings. 5.3. EEG coherence and WPLI networks We see there is a large difference in the integration, modularity and complexity of the EEG coherence and dW- PLI networks (Fig. 3, top left, centre left and centre right, respectively). The EEG coherence networks (blue shade) behave similarly to the ring (blue lines) and grid lattice (yellow lines) networks, agreeing with the volume conduc- tion effects that dominate zero-lag dependency measures [43], i.e. the closer the nodes are the stronger the weights are. The dWPLI networks (red shade) on the other hand have a more integrated and less modular nature, which re- flects the notion that phase-based functionality mitigates volume conduction effects and is thus less confined by anatomical structure [43]. The very high complexity of the dWPLI networks (and very possibly phase-lag measures in general [47]) provides evidence to support that phase-based connectivity does indeed largely overcome the volume conduction effect and therefore maintains a richer complexity echoing the com- plex interactions of brain functionality [1]. With regards to how the WCH model advances our understanding of dWPLI and coherence network differ- ences, we note that the high segregation of the coherence networks (Fig. 3 top left) is approached by the WCH model with high values of strength parameter (Fig. 5, top left) and is comparable with regular lattice and grid lat- tice CWN curves (Fig. 3, top left, blue and yellow lines, respectively), denoting a move to a more strict class-based Figure 7: Clustering coefficient, C, degree variance, V , modularity, Q, and complexity, R, against connection density, P , of binarised weighted networks, for WCH model (red, median (m) ± interquartile range (iqr)) and EEG dWPLI neworks (blue, m ± iqr). not driven by integration, arising as a middle ground be- tween regular and random systems as previously conjec- tured [7, 9], but, driven by hierarchical complexity, arising in the middle ground between weak hierarchical topology or 'all nodes are equal' systems, such as random or regular networks, and strong hierarchical topology, such as star or strict class-based systems including grid lattice and frac- tal modular networks (see Fig. 5). Thus the hierarchical structure can be seen as a key aspect of the complexity inherent in complex systems. Impressively, the dWPLI EEG networks display a gen- erally greater hierarchical complexity than that expressed by our model which is specifically designed to probe com- plex interactions in hierarchical structures. Thus we pose such complexity as a key aspect of brain function as mod- elled by phase-based connectivity. 5.2. Weighted complex hierarchy as null model There are two clear reasons why the WCH model is a good fit for functional connectivity networks from EEG recordings. Not only does it create several hub like nodes giving a high degree variability, but furthermore it sim- ulates the rich club phenomena found in complex brain networks [22, 44], as the higher the hierarchy levels of two nodes, the stronger the weight of the connection will be between them, see Fig. 1.C. One of the greatest benefits of this model over others is that it simulates brain networks previous to network pro- cessing steps because it creates CWNs rather than sparse networks. This means that any and all techniques one wants to use on the brain networks can be applied el- egantly and in parallel with this single null model free from any complications. Particularly, methods which cre- ate sparse binary networks directly, whether these models 9 topology. This is also reflected in the hierarchical com- plexity (bottom right of corresponding figures), where the lower complexity peaking at a later density to dWPLI (Fig. 3, centre right) is mimicked in the behaviour of increasing strength parameter in the WCH model (Fig. 5, bottom right). This provides further evidence of the relevance and flexibility of the WCH model. In contrast there is an evi- dent lack of ability to make similar comments with respect to the popular small world and scale-free models. This criticism can be extended towards network models which randomise connections while maintaining degree distribu- tions [18], since such an enforced topological attribute does not allow one to analyse how that very important attribute is actually constructed. Future work will provide extension to modular structures in our model to focus on what roles modularity plays on these aspects, since Q and V behave contrastingly to this extrapolation. 5.4. Dense scale-free networks A striking feature seen is in the degree variance curves where a highly symmetric parabolic curve is noted with a central maximum value for random graphs, WCH net- works and EEG networks. This feature reveals to us a 'scale-free' paradigm at all density levels and not just the classic sparse network scale-free at low densities. In other words, the scale-free nature found in brain networks is first and foremost encoded in the connectivity weights, which, through selective binarisation, therefore can reveal to us the scale-free property as expressed at different density ranges. As the density of the network increases one obtains more even distributions of high and low density nodes, in- dicated by the high values of V , and, eventually, towards high densities the symmetry of V values with low densi- ties tells us that the scale-free network is characterised by a small number of low degree nodes and a majority of high degree nodes, i.e. the inverse (or complement) of the low density behaviour. 5.5. Topological randomness If we define a uniformly random topology as that which exhibits a uniform distribution of topological values over the space enveloped by the minimal and maximal topolo- gies, it is very apparent that E-R random networks do not satisfy this criteria, but, instead, have a restricted topol- ogy at all density levels where the interquartile range is much smaller in comparison with that of the EEG net- works and the proposed null model. We thus see that uniformly distributed random weights do not lead to a uniformly random topology in this sense, but instead to a very particular optimally integrated, moderately regular, lowly modular and low complexity topology at all densi- ties. Based on this evidence and previous discussion of random networks in the methods section, we suggest that E-R random networks should be re-understood as opti- mally integrated networks. Following from this the randomisation of connections used widely in null models is not a topologically randomis- ing process but, more accurately, a topologically integra- tive process. Such a feature is then not necessarily typical of network topology and thus one must be cautious to use this as a null model unless one wants to specifically target integrative behaviour. Further, the practice of normalisa- tion of graph values by E-R random graph values [48] must also be used with due caution. The basis of such a normal- isation is to contrast a networks values with those of the 'average' network topology [49], rather than contrasting with a highly specific topology which behaves very differ- ently to real world networks [50]. This evidence provides further justification for the adoption of our WCH model as a relevant and powerful replacement to these models. 6. Conclusion We introduced a metric for measuring the hierarchical complexity of a network and a highly flexible and elegant WCH model. These provided key insights into what distin- guishes functional brain networks from both ordered and spontaneous forms as generally the most complex kind of topology and the important role that hierarchical struc- ture plays in this. Further, we showed that phase-based connectivity topology was more complex than amplitude influenced connectivity topology, which we extrapolated as due to the more ordered structure enforced by volume conduction effects. In our analysis we constructed a frame- work for CWNs for brain functional connectivity to replace the framework for sparse networks adopted from other net- work science research areas. This included the synthesis of concepts from the literature in a succinct manner and the generalisation of sparse binary archetypes to CWN form. The perspective allowed by this comprehensive analysis provided new evidence regarding key factors of network topology in general. Importantly we provided evidence of the non-topologically random nature of uniformly random weighted networks. From this it follows that our model is more relevant and appropriate than prevalent connection randomisation processes. Also, a scale-free paradigm was extended to all network densities. Particularly, these in- sights help towards a comprehensive understanding of the framework within which functional connectivity networks are set and thus provide invaluable information and tools for future clinical and technical research in neuroscience. Matlab codes for all synthesis and analysis of the networks as introduced in this paper are publicly available on pub- lication at http://dx.doi.org/10.7488/ds/1520. 7. Acknowledgements Keith Smith is funded by the Engineering and Physical Sciences Research Council (EPSRC). 10 References [1] E. Bullmore, O. Sporns, "Complex brain networks: graph the- oretical analysis of structural and functional systems", Nature Neuroscience Review, 10: 186-198, 2009. [2] C.J. Stam, "Modern network science of neurological disorders", Nature Neuroscience Review, 15: 683-695, 2014. [3] D. Papo, J. M. Buldu, S. Boccaletti, E. T. Bullmore, "Complex network theory and the brain", Phil. Trans. R. Soc. B, 369: 20130520, 2014. [4] E. Ravasz and A.L. Barab´asi, "Hierarchical organization in com- plex networks", Phys. Rev. E, 67: 026112, 2003. [5] D. Meunier, R. Lambiotte & E. T. Bullmore, "Modular and hi- erarchically modular organisation of brain networks", Frontiers in Neurscience, doi: 10.3389/fnins.2010.00200, 2010. [6] M. Costa, A.L. Goldberger, C.-K. Peng, "Multiscale entropy analysis of biological signals", Phys. Rev. E, 71: 021906, 2005. [7] G. Tononi, O. Sporns, G.M. Edelman, "A measure for brain complexity: Relating functional segregation and integration in the nervous system", Proc Natl Acad Sci, 91(11):5033-7, 1994. [8] R.Sol´e & S. Valverde, "Information theory of complex networks: on evolution and architectural constraints", Lect. Notes Phys., 650: 189-207, 2004. [9] D.J. Watts & S.H. Strogatz, "Collective dynamics of small- world networks",Letters to Nature, 393: 440-442, 1998. [10] O. Sporns, "Small-world connectivity, motif composition, and complexity of fractal neuronal connections", BioSystems, 85: 5564, 2006. [11] M. Rubinov, O. Sporns, "Weight-conserving characterization of complex functional brain networks", Neuroimage, 56(4):2068- 2079, 2011. [12] F. D.V. Fallani, J. Richiardi, M. Chavez, S. Achard, "Graph analysis of functional brain networks: practical issues in trans- lational neuroscience", Phil. Trans. R. Soc. B: Biological Sci- ences, 369 (1653): 20130521, 2014. [13] C. Li, H. Wang, W. de Haan, C.J. Stam, P. Van Mieghem, "The correlation of metrics in complex networks with applications in functional brain networks", J. Stat. Mech., doi:10.1088/1742- 5468/ 2011/11/P11018, 2011. [14] P. Tewarie, E. van Dellen, A. Hillebrand, C.J. Stam, "The min- imum spanning tree: an unbiased method for brain network analysis", Neuroimage, 104: 177-188, 2015. [15] A.L. Barabasi & R. Albert, "Emergence of Scaling in Random Networks", Science, 286: 509-512,1999. [16] M.E.J. Newman, "Networks", Oxford University Press, Oxford, 2010. [17] O. Sporns, "Networks of the Brain", MIT Press, MA, 2010. [18] M. Rubinov, O. Sporns, "Complex network measures of brain connectivity: Uses and interpretations", NeuroImage, 52:1059- 1069, 2010. [19] V.M. Eguiluz, D.R. Chialvo, G.A. Cecchi, M. Baliki, A. V. Ap- karian, "Scale Free Brain Functional Networks", Physical Re- view Letters, 94: 018102, 2005. [20] T.A.B. Snijders, "The degree variance: an index of graph het- erogeneity", Social Networks, 3(3): 163-174, 1981. [21] A. Sandryhaila, J.M.F. Moura, "Discrete Signal Processing on Graphs", IEEE Transactions on Signal Processing, 61(7):1644- 1656, 2013. [22] J.J. McAuley, L.F. Costa, T.S. Caetano, "The rich-club phe- nomena across complex network hierarchies", Appl. Phys. Lett., 91, doi: 10.1063/1.2773951, 2007. [23] P. Erdos & A. R´enyi, "On Random Graphs", Publicationes Mathematicae Debrecen, 6:290-297, 1959. [24] P. Bonacich, P. Lloyd, "Eigenvector-like measures of central- ity for asymmetric relations", Social Networks, 23(3): 191-201, 2001. [25] M. Molloy & B. Reed, "A critical point for random graphs with a given degree sequence", Random Structures & Algorithms, 6 (2-3): 161-180, 1995. [26] S. Milgram, "The Small World Problem", Psychology To- day,1(1): 61-67, 1967. [27] E. Bullmore, O. Sporns, "The economy of brain network organ- isation", Nature, 13:336-349, 2012. [28] A.L. Barabasi, R. Albert & H. Jeong, "Mean-field theory for scale-free random networks", Physica A, 272(1-2): 173-187, 1999. [29] B.M. Tijms, A.M. Wink, W. de Haan, W.M. van der Flier, C.J. Stam, P. Scheltens, F. Barkhof, "Alzheimer's disease: connect- ing findings from graph theoretical studies of brain networks", Neurobiology of Ageing, 34: 2023-2036, 2013. [30] J. Petersen, "Die Theorie der regularen Graphs", Acta Math., 15: 193-220, 1891. [31] B.M. ´Abrego, S. Fernndez-Merchant, M.G. Neubauer, W. Watkins, "Sum of squares of degrees in a graph", Journal of Inequalities in Pure and Applied Mathematics, 10(3): 64, 2009. [32] C.J. Stam, B.F. Jones, G. Nolte, M. Breakspear, P. Schel- tens, "Small-World Networks and Functional Connectivity in Alzheimer's Disease", Cerebral Cortex, 17:92-99, 2007. [33] M. Lynall, D. S. Bassett, R. Kerwin, P.J. McKenna, M. Kitzbichler, U. Muller & E. Bullmore, "Functional Connectiv- ity and brain networks in schizophrenia", J Neurosci, 30(28): 9477-9487, 2010. [34] C.E. Ginestet, T.E. Nichols, E.T. Bullmore, A. Simmons, "Brain network analysis: separating cost from topology us- ing cost-integration", PloS One, 6, e21570, doi: 10.1371/jour- nal.pone.0021570, 2011. [35] A.L. Goldberger, L.A.N. Amaral, L. Glass, J.M. Hausdorff, P.C. Ivanov, R.G. Mark, J.E. Mietus, G.B. Moody, C.K. Peng, H.E. Stanley, "PhysioBank, PhysioToolkit, and Phys- ioNet: Components of a new research resource for complex physiologic signals", Circulation, 101(23): e215-e220, dataset: doi:10.13026/C28G6P, 2000. [36] M.E.J. Newman, M. Girvan, "Finding and evaluating commu- nity structure in networks", Phys. Rev. E, 69(2): 026113, 2004. [37] M.E.J. Newman, "Modularity and community structure in net- works", Phys Rev E, 23: 8577-8582, 2006. [38] V.D. Blondel, J-L. Guillaume, R. Lambiotte, E. Lefebvre, "Fast unfolding of communities in large networks", J. Stat. Mech., doi: 10.1088/1742-5468/2008/10/P10008, 2008. [39] C.E. Shannon, "A mathematical theory of communication", Bell System Technical Journal, 27: 623656, 1948. [40] G. Schalk, D.J. McFarland, T. Hinterberger, N. Birbaumer, J.R. Wolpaw, "BCI2000: A General-Purpose Brain-Computer Inter- face (BCI) System", IEEE Transactions on Biomedical Engi- neering, 51(6):1034-1043, 2004. [41] R. Oostenveld, P. Fries, E. Maris and J-M. Schoffelen, "Field- Trip: Open Source Software for Advanced Analysis of MEG, EEG, and Invasive Electrophysiological Data", Computational Intelligence and Neuroscience, Volume 2011, 156869, 9 pages, 2011. [42] M. Vinck, R. Oostenveld, M. van Wingerden, F. Battaglia, C.M.A. Pennartz, "An improved index of phase-synchronization for electrophysiological data in the presence of volume- conduction, noise and sample-size bias", NeuroImage, 55:1548- 1565, 2011. [43] E. van Diessen, T. Numan, E. van Dellen, A.W. van der Kooi, M. Boersma, D. Hofman, R. van Lutterveld, B.W. van Dijk, E.C.W. van Straaten, A. Hillebrand, C.J. Stam, "Opportunities and methodological challenges in EEG and MEG resting state functional brain network research", Clinical Neurophysiology, doi:10.1016/j.clinph.2014.11.018, 2014. [44] M.P. van den Heuvel, O. Sporns, "Rich-Club Organisation of the Human Connectome", Journal of Neuroscience, 31(44): 15775- 15786, 2011. [45] C.J. Stam, P. Tewarie, E. Van Dellen, E.C.W. van Straaten, A. Hillebrand, P. Van Mieghem, "The trees and the forest: Char- acterization of complex brain networks with minimum spanning trees", International Journal of Psychophysiology, 92: 129-138, 2014. [46] K. Smith, H. Azami, J. Escudero, M.A. Parra, J.M. Starr, "Comparison of Network Analysis Approaches on EEG Con- nectivity in Beta during Visual Short-Term Memory Tasks", 11 Proceedings of the IEEE EMBC15, 2207 - 2210, 2015. [47] J. Dauwels, F. Viallate, T. Musha, A. Cichocki, "A compar- ative study of synchrony measures for the early diagnosis of Alzheimer's disease based on EEG", NeuroImage, 49(1): 668- 693, 2010. [48] M.D. Humphries, K. Gurney, "Network 'Small-world-ness': A Quantitative Method for Determining Canonical Network Equivalence", PLoS One, 3(4): e0002051, 2008. [49] B. Bollob´as, "Random Graphs", ch.8 of "Modern Graph The- ory", Graduate Texts in Mathematics, Springer New York, 184: 215-252, doi:0.1007/978-1-4612-0619-4 7, 1998. [50] M.E.J. Newman, "Random graphs as models of networks", ch.2 from S. Bornholdt, H.G. Schuster,"Handbook of Graphs and Networks: From the Genome to the Internet", Wiley, UK, DOI: 10.1002/3527602755, 2006. 12
1106.2048
1
1106
2011-06-09T18:30:52
Old equations for a new system: A possible use of Navier-Stokes equations to model the circulation of spikes in the nervous system
[ "q-bio.NC" ]
In the present work we discuss a possible application of Navier-Stokes-based models to the quantitative description of the circulation of nervous impulses throughout the nervous system. In previous works we have shown that the discharge from Basal Ganglia neurons from patients with Parkinson's disease share mathematical features with the velocity fields of fluids under turbulence regimes. In the present work we try to build the fundaments for a physical analogy between both kinds of systems.
q-bio.NC
q-bio
Old equations for a new system: A possible use of Navier-Stokes equations to model the circulation of spikes in the nervous system Daniela Sabrina Andres Department of Physiology, Medicine School, University of Buenos Aires CONICET Institute for Neurological Research Raúl Carrea, Movement Disorders Section, Neuroscience Department, FLENI Montañeses 2325, C1428AQK, Buenos Aires, Argentina [email protected] Abstract In the present work we discuss a possible application of Navier-Stokes-based models to the quantitative description of the circulation of nervous impulses throughout the nervous system. Velocity and momentum of spikes are not usually considered when modeling information transmission in the nervous system, however in previous works we have shown that the discharge from Basal Ganglia neurons from patients with Parkinson’s disease share mathematical features with the velocity fields of fluids under turbulence regimes. In particular the properties that are similar are related to the behavior of energy, and therefore fluid dynamics’ concepts might be useful in the future for the understanding of the nervous system from a mathematical perspective. In the present work we try to build the fundaments for a physical analogy between both kinds of systems. Introduction To present date, the structure of the neural code, or even if there is one, remains unknown.1,2 Several approaches have been proposed to solve the riddle. Different works have discussed the idea of information being coded in a rate code or a time code.3,4 In the first case the exact time of occurrence of a single spike could be random, while in the second case a deterministic time pattern is thought to code for information. In a recent work of our group, it has been proposed that information could also be coded in a scale code, with different scales of the neuronal discharge transmitting either complementary or redundant information.5 Of course, different combinations of the previous might be possible, and the universality of the code is also under discussion. 1 Different neural systems might use different codes and even switch to different forms of coding under different circumstances. 6 ,7 In the present work we will discuss a general frame to analyze the circulation of nervous impulses throughout the nervous system. In previous works we have shown that properties measured from the brain from patients with Parkinson’s disease (PD) have similarities with turbulent fluids.5,8,9 We have proposed that Navier- Stokes-like equations might be useful for the modeling of the in transmission information central nervous system. In the present paper we present a preliminary description of a physical analogy and a possible way in which the formalism of Navier-Stokes (NS) might apply for the study of information transmission in a neural system. A brief discussion about the possible consequences of the present work might be appropriate. In clinical neurology the question about the way in which neurons code information the understanding and becomes relevant for modeling of information transmission in the central nervous system. In particular in the case of Parkinson’s disease, the application of Deep Brain Stimulation therapy (DBS) affects the transmission of nervous impulses through the basal ganglia- loop.10,11 DBS thalamic-cortical (BG-Th-Ctx) consists of chronically implanting a stimulation electrode in a chosen neural center and it has proven to succeed in the treatment of different conditions.12 However, to present date the mechanisms of action of DBS therapy are not fully understood.13 ,14 This poor understanding of the exact mechanisms of action of DBS has important practical consequences. Today, there exist limitations in the outcome prediction of patients with PD who are treated with DBS, and the indications of this therapy are limited to the most severe cases.15 Therefore the fact that we currently lack a mathematical model of DBS poses important limitations to this treatment. Based on some experimental evidence presented in previous works 5 ,8,9 we will discuss a possible analogy between the circulation of spikes in the BG-Th-Ctx loop and the dynamics of fluids. Fundaments of the model The present model is centered on the description of the velocity of transmission of the nervous impulse in the nervous system. Although velocity is frequently not used to describe the behavior of neural systems, it might possibly be a crucial variable for the understanding of transmission.16 To understand information the situation that we are trying to represent, imagine a simple neural system like the one in figure 1. Suppose that a spatially homogeneous stimulus S1 is applied to three adjacent neurons, N1-3, forming layer one (LI) of a neural network. All three neurons receive the stimulus at the same time, called ti, and we propose that the three of them will respond with the generation of a single spike to the stimulus. The problem that concerns us is at what times, called t1, t2 and t3, the spikes originated by S1 will arrive to the next layer of the network travelling through the different neurons. (For simplification we will consider here that the next layer lies at a constant distance from layer one, although the spatial architecture of physiological neural systems adds a complexity level to the problem that we are analyzing. This fact could be modeled with the use of non-Bravais lattices, i.e. integration meshes without translational symmetry.) Assuming that spikes travel at the same velocity through all the axons of the layer, t1, t2 and t3 will depend on the time consumed in summation and therefore on the level of excitation of each neuron. We assume that all the cells in a layer have the same time and space constants and firing threshold. Therefore changes in membrane potential will occur at the same rate in each cell when the depolarizing stimulus is applied and, if the stimulus is enough to produce a supra- Figure 1) Although the velocity of spikes is a quantity that is not commonly discussed, it might well describe the circulation of impulses in the nervous system. Suppose that three adjacent neurons are affected simultaneously by a spatially homogeneous stimulus at an initial time t, and assume that all three neurons have the same time and space constants and firing threshold and a constant axonal length. If the three neurons respond with the generation of a single spike to the stimulus, then the times t1, t2 and t3 at which the generated spikes will arrive to the next neuronal layer do not need to be the same, but will depend on the exact level of excitation at of each neuron at time t. 2 threshold depolarization, then the time at which the response will be generated depends on how far the membrane potential lies from the threshold at the moment the cell received the input. It can be easily seen that this difference in the initial excitation level of each neuron will translate into a difference between t1, t2 and t3, a phenomenon that can actually be interpreted as a change of velocity of the impulse if we consider the network globally. This acceleration effect could be potentiated or attenuated in the next layer, adding variability to the velocity of transmission of spikes in the nervous system and finally generating the complex interspike intervals time series that are usually measured in neural systems. So far we have described a simple fact about neural transmission that is often not taken into account. This is also true for another simple fact that needs to be considered prior to the discussion of the present model. Given two neurons N1 and N2, any of them will have a greater probability of applying a positive acceleration to an impulse I if the other one is doing so. This is equivalent to saying that the velocity through a given neuron will be minor if the adjacent neuron is not applying any acceleration to the impulse. This can be explained with our current knowledge of extracellular modulation of synaptic activity, a function conducted primarily by astrocytes. 17-19 The importance of the finite extracellular space and its chemical composition in neural conduction is being acknowledged progressively, and is a possible mechanism of coupling between adjacent cells. 20,21 When considering a large network of neurons this tendency of adjacent neurons to show similar acceleration values will have the effect of helping a spatially homogeneous stimulus to remain so. In other words, while acceleration and deformation in time and space of the stimulus travelling through the network is possible, the stimulus will essentially resist to this deformation. We will build an analogy between this fact and the viscosity of fluids, and will call this property of stimuli travelling through large ensembles of neurons, µ . Having stated the velocity of that transmission of an impulse from one layer to the next one is not necessarily constant and that a spatially homogeneous stimulus naturally resists deformation, we have the bases needed to apply the present NS-based model to the transport of nervous stimuli through the neural network that we are considering. We can write the well known Navier- Stokes equations as follows: P µ ∇+ 2 r v (2) (1) −∇= r = 0 g r ∂⇒ v ρ ∂ t Where µ is the viscosity, P stands for pressure and v for velocity. We will consider g = 0 in order to ignore the gravity term. To make use of this equation for an interpretation of the movement of nervous impulses in the nervous systems, we will define ρ as the density of impulses I in a given volume of nervous tissue. Pressure can also be considered as the energy of the system per unit of volume. For that reason the best analogy possible in the nervous system should be a function of the excitation level of the neuron. In time, the level of excitation can be described as a function of the sum of synaptic weights to make in order comparisons with other models. Since the general state of excitation depends on the sum of synaptic weights of all the synapses present in the system, we can write excitation, e, as a function of this sum:  =1i = ∑  eP  m iw    (3) What was said to this point allows us to state that the propagation of the nervous impulse through the nervous system requires energy per volume unit. We have already stated that µ will account for the resistance to deformation of a spatially homogeneous stimulus. These ideas are just being formulated in a conceptual way in the present work, and experimental research needs to follow in order to describe in a quantitative way these properties. A few words need to be said in reference to a possible comparison to experimental data. Interspike intervals (ISIS) are a common, robust and reliable measure used for the analysis of experimental data in neural systems. Since the detection of spikes is a robust procedure itself, these time series have low levels of noise, making them suitable for the analysis of non-linear properties. Interspike intervals time series are 3 constructed using the differences between the times of appearance of spikes, and are therefore a measure of time. A practical difficulty that we will encounter when trying to apply NS-based model to the circulation of impulses in the nervous system is that most works based on NS models analyze velocity fields. For that reason a way to compare time and velocity would be desirable, and some transformation between both variables might be needed. This transformation can be performed making use of single cell recordings. When analyzing single unit data, it can be considered that every spike is generated at the same point of space. This allows us to take distance as 1, and so a relationship of x-1 relates both variables. Of course, some normalization would be needed when working with velocity fields in more than one dimension. Then the inverse relationship would be to the norm of vector velocity, and the relationship would be: ),( txI ISI v = −= t i ∆ x ∆ t    , t + 1 i ∆ y ∆ t , ∆ z ∆ t    (4) (5) (6) ∆ x LI − LII ∆= y LI − LII ∆= z LII LI − = 1 (7) ⇒ v LI − LII = 3 ISI (8) Here I(x,t) describes the impulse as a function of time and space and ∆xLI-LII, ∆yLI-LII and ∆zLI-LII refer to the distance in three dimensional space between neuronal layers I and II (we are essentially considering a cube of side =1). Discussion During the last decades our understanding of BG physiology and pathophysiology has been challenged by the introduction of DBS therapy. 22- 23 The observations made in the context of stimulation treatment of human patients with PD as well as functional neurosurgery showed many inconsistencies and paradoxes with the classical Albin-De Long’s model of the BG, opening the way for new models, different than this classical “box and arrow” view.24 There are many important points that the classical model fails to account for. First of all, it has been shown that high frequency DBS of the GPi is effective in the treatment of 4 hyperkinetic (chorea) as well as hypokinetic (PD) disorders in exactly the same way. 25-26 This fact is contradictory with the notion that hyperkinetic and hypokinetic diseases are based on opposite pathogenic mechanisms. Secondly, stimulation of almost any nuclei of the BG-Th-Ctx loop yields positive results in the treatment of at least some PD symptoms. This includes stimulation of GPi, GPe, ventrolateral Thalamus, subthalamic nucleus, motor cortex and even zona incerta. 27-31 Finally, it has been described that HF stimulation of the GPi increases the activity of this nucleus, which is exactly the mechanism proposed as the origin of the symptoms in the classical model of PD. 32-35 In this context, new hypothesis have been made about the functioning of the normal and diseased BG. Current models include dimensionality reduction models, action-selection devices, learning reinforcement and some neural network models. 36- 40 These models capture some properties of the dynamics, but up to now none of them has proven successful enough to accurately predict outcome and allow a better programming of the devices. It has also been proposed that non-linear features might be the basis of the functioning of the BG- Th-Ctx loop and that some of these properties might be altered in PD. 41-45 However, currently there is not a mathematical model that allows representing and analyzing these properties in a quantitative fashion. In previous works we have found some interesting mathematical similarities between the ISIS time series of parkinsonian GPi neurons and velocity fields of fluids undergoing turbulence. 5 ,8,9 In particular, an exponential decay of the power spectrum makes it possible to talk about critical scale length, a property that is not shared by many systems in the nature. This form of decay of the spectrum is typical from turbulent fluids, and follows an analytic solution of Navier-Stokes equations when energy remains bounded. 46,47 Therefore it could be thought that energy behaves in a similar manner in both systems. Other similarities that we have reported refer to the existence of positive Lyapunov exponents 48 ,49 and a scaling of the structure function that is typical from turbulent fluids too 50 and that lead to the description of a multifractal behavior in the signals studied. In light of the evidence found up to this point it seems reasonable to speculate that the equations underlying the dynamics of the neural system under discussion might be similar in form to those that better describe the behavior of fluids: Navier-Stokes equations. A further step is what we have tried in the present work, and it consists of looking for a physical analogy that explains the sense in which each term of the equations might apply for neural systems. Another type of neurophysiological model are neural fields. These are concerned with firing rates more than with velocity, 51-55 although sometimes they have been used to describe the changes of velocity present in the nervous system16. However they are mainly based on reaction-diffusion kinds of systems and work under the assumption of density heterogeneities, a description that lacks of a strong physiological basis. In turn these heterogeneities are usually modeled as continuous functions of space, a condition that does not seem very physiological either. So we can say confidently that up to now there is not a model that handles the behavior of the velocity of spikes in the nervous system in a quantitative and physiological enough manner. Currently there are several experimental approaches to the study of PD and DBS. Under some experimental settings, neurons from the basal ganglia from patients with PD have shown an oscillatory behavior.56-59 Under this evidence it has been proposed that a pathologic increment of oscillatory activity or excessive synchronization between neurons might be the pathogenic mechanism of PD.60 However, other works have shown the mentioned neurons to behave in a complex fashion, with an irregular output.8,43 The fundamental problem is to what extent these different works are comparable. Probably at least some of the mathematical properties of the signals studied depend on the experimental settings used, such as in vitro vs. in vivo experiments, the level of external input of the neuron, the type of anesthesia used, etc. 61-65 In the previous works by our group that we have mentioned before, we have worked with data from GPi neurons form patients with PD. This data were recorded during functional neurosurgery, with the patients awake and under local anesthesia only. Because of that reason the findings observed might be due at least to some extent to the state of consciousness of the patients, which defines the high dimensional input from the neurons studied at the times of the recordings. Since the mathematical properties 5 analyzed were described under these particular experimental conditions, some questions follow that will need to be considered in future works. A control group of a healthy population of neurons will be needed. For the reasons that were discussed, the best control group would probably consist of recordings performed on animals awake with microelectrode implanting techniques. This experiment together with the analysis of anesthetized animals could help elucidate if the observations under discussion are due to the presence of PD itself or to the state of consciousness of the animal and to what extent. Of course, the hypothesis presented here need to be tested. Many directions of research follow from the idea that NS-based models might be useful for the modeling of the nervous system. First of all, an analytic work might be useful to test the relation between ISIS and velocity fields, i.e. to test if the results already known for velocity fields and the properties observed for ISIS time series hold under an x-1 transformation. Secondly, to perform simulations for simple conditions might be useful, obtaining a comparable measure from them (times between events). For example, tissue infarctations from different sizes and shapes might be a good starting point to run simulations (different geometric boundary conditions). To model the effect of DBS it would be probably necessary to work with the pressure term of the equations. Since DBS stimulates the nervous tissue with a pulse of electric tension, it can be interpreted that it produces a change in the level of energy of the tissue (by affecting ionic gradients for example), and an analogy could be easily built with the pressure of fluids. Finally, new experimental approaches could be design to study the movement of impulses in nervous tissue from the perspective of velocity. However, and although new experiments might be desirable, it needs to be considered that experimental complexity needs to be kept to a minimum when working with human patients. For that reasons new experiments might be necessarily conducted in laboratory animals, or even more brain slices or cell cultures and, again, comparison with data from human brains will become a problem to be considered. Finally, if the presented approach turns out to be useful for the modeling of the BG, the universality of the model will become a question, and other neural systems under different conditions (awake, anesthetized, in vivo vs. in vitro) will need to be tested. 8. Andres DS, Cerquetti D, Merello M. Finite dimensional structure of the GPi discharge in patients with Parkinson’s Conclusion In the present work we do not present simulations or computational results. We have explained the fundaments of a possible application of Navier-Stokes-based equations for the modeling of the movement of nervous impulse in the nervous system. Velocity and momentum of spikes are not usually considered when modeling information transmission in the nervous system, but recently we have found evidence that interspike time series share some intervals mathematical properties with the behavior of velocity fields of fluids under turbulence. In particular these properties are related to the behavior of energy, and therefore fluid dynamics’ concepts might be useful in the future for the understanding of the nervous system, both from a mathematical perspective and helping us to find new physical analogies. References 1. Shadlen MN, NewsomeWT. Noise, neural codes and cortical organization. Curr Opin Neurobiol 1994; 4(4):569– 79. 2. Stein RB, Roderich Gossen ER, Jones KE. Neuronal variability: noise or part of the signal? Nat Rev 2005; 6:389–97. disease. Int J Neural Syst 2011; in press. 9. Andres DS, Bocaccio H, Cerquetti DF, Merello M. On the multiple scale of the neural code. 2011; under review. 10. Hashimoto T, Elder CM, Okun MS, Patrick SK, Vitek JL. Stimulation of the subthalamic nucleus changes the firing pattern of pallidal neurons. J of Neurosc 2003; 23(5): 1916 –23. 11. Lanotte MM, Rizzone M, Bergamasco B, Faccani G, Melcarne A, Lopiano L. Deep brain stimulation of the subthalamic nucleus: anatomical, neurophysiological and outcome correlations with the effects of stimulation. J Neurol Neurosurg Psychiatry 2002; 72: 53-58. 12. Pereira EAC, Green AL, Nandi D, Aziz TZ. Deep Brain Stimulation: indications and evidence. Expert Reviews 2007; 4(5): 591-603. 13. Montgomery Jr. EB, Gale JT. Mechanisms of action of Deep Brain Stimulation (DBS). Neuroscience and Biobehavioral Reviews 2008; 32: 388–407. 14. Nambu A. Seven problems on the basal ganglia. Curr Opin Neurob 2008; 18(6): 595–604. 15. Rodriguez-Oroz MC, Obeso JA, Lang AE, Houeto JL, Pollak P, et al. Bilateral deep brain stimulation in Parkinson’s disease: a multicentre study with 4 years follow-up. Brain 2005; 128: 2240-49. 16. Bressloff PC. Traveling fronts and wave propagation failure in an inhomogeneous neural network. Physica D (2001); 155: 83-100. 17. Newman EA. New roles for astrocytes: regulation of 3. Bialek W, Rieke F, de Ruyter van Steveninck RR, Warland synaptic transmission. Trends in Neurosci 2003; 26(10): D. Reading a neural code. Science 1991; 252(5014):1854- 57. 4. Sejnowski TJ. Time for a new neural code? Nature 1995; 376:21. 5. Andres DS, Cerquetti D, Merello M. Turbulence in Globus Pallidum neurons in patients with Parkinson’s disease: Exponential decay of the power spectrum. J Neurosci Meth 2011; 197(1): 14–20. 6. Mehta MR, Lee AK, Wilson MA. Role of experience and oscillations in transforming a rate code in a temporal code. Nature 2002; 417:741-46. 536-42. 18. Araque A, Carmignoto G, Haydon PG. Dynamic signaling between astrocytes and neurons. Ann Rev Phys 2001; 63: 795-813. 19. Fellin T. Communication between neurons and astrocytes: relevance to the modulation of synaptic and network activity. Jof Neuroch 2009; 108(3): 533-44. 20. Gloor SM. Relevance of NA,K-ATPase to local extracellular potassium homeostasis and modulation of synaptic transmission. FEBS Letters 1997; 412(1):1-4. 21. Fields RD, Stevens B. ATP: an extracellular signaling molecule between neurons and glia. Trends in Neurosci 7. Fairhall AL, Lewen GD, Bialek W, de Ruyter van 2000; 23(12):625-33. Steveninck RR. Eficiency and ambiguity in an adaptive neural code. Nature 2001; 412:787-92. 6 22. Wichmann T, DeLong MR. Deep brain stimulation for 34. Albin RL, Young A, Penny JB. The functions anatomy of neurologic and neuropsychiatric disorders. Neuron 2006; basal ganglia disorders. Trends in Neurosci 1989; 12: 366- 52: 197–204. 75. 23. Sestini S, Ramat S, Formiconi AR, Ammannati F, Sorbi S, 35. DeLong MR. Primate models of movement disorders of Pupi A. Brain Networks Underlying the Clinical Effects of basal ganglia origin. Trends in Neurosci 1990; 13: 281–85. Long-Term Subthalamic Stimulation for Parkinson’s Disease: A 4-Year Follow-up Study with rCBF SPECT. J Nucl Med 2005; 46:1444–54. 24. Bar-Gad I, Bergman H. Stepping out of the box: information processing in the neural networks of the basal ganglia. Current Opinion in Neurobiology 2001; 11: 689– 95. 25. Montgomery Jr. EB. Deep brain stimulation for hyperkinetic disorders. Neurosurgical Focus 2004; 17: E1. 36. Berns GS, Sejnowski TJ. A Computational Model of How the Basal Ganglia Produce Sequences. Journal of Cognitive Neuroscience 1998; 10(1): 108-121. 37. Gurney KN, Prescott TJ, Redgrave P. The Basal Ganglia viewed as an Action Selection Device. Proceed of the Eighth Int Conf on Artif Neural Net, September 2-4 1998, Skövde, Sweden, 1033–38. 38. Terman D, Rubin JE,. Yew AC, Wilson CJ. Activity Patterns in a Model for the Subthalamopallidal Network of 26. Montgomery Jr. EB. Dynamically coupled, high-frequency the Basal Ganglia. J of Neurosci 2002; 22(7): 2963–76. reentrant, non-linear oscillators embedded in scale-free basal ganglia-thalamic-cortical networks mediating function and deep brain stimulation effects. Nonlinear Studies 2004; 11: 385–421. 27. Deep Brain Stimulation in Parkinson’s Disease Group. Deep-brain stimulation of the subthalamic nucleus or the pars interna of the globus pallidus in Parkinson’s disease. 39. Bar-Gad I, Morris G, Bergman H. Information processing, dimensionality reduction and reinforcement learning in the basal ganglia. Prog in Neurobiol 2003; 71: 439–473. 40. Doya K. What are the computations of the cerebellum, the basal ganglia and the cerebreal cortex? Neural Networks 1999; 12: 961-9. New England and Journal of Medicine 2001; 345: 956–63. 41. Darbin O, Soares J, Wichmann T. Nonlinear analysis of 28. Koller W, Pahwa R, Busenbark K, Hubble J, Wilkinson S, et al. High-frequency unilateral thalamic stimulation in the discharge patterns in monkey basal ganglia. Brain Res 2006; 1118: 84–93. treatment of essential and parkinsonian tremor. Ann of 42. Lim J, Sanghera MK, Darbin O, Stewart RM, Jankovic J, Neurol 1997; 42: 292–99. 29. Vitek JL, Hashimoto T, Peoples J, DeLong MR, Bakay AE. Acute stimulation in the external segment of the Simpson R. Nonlinear temporal organization of neuronal discharge in the basal ganglia of Parkinson’s disease patients. Exp Neurol 2010; 224(2): 542–44. globus pallidus improves Parkinsonian motor signs. 43. Obeso JA, Rodríguez-Oroz MC, Rodríguez M, Lanciego Movement Disorders 2004; 19: 907–15. JL, et al. Pathophysiology of the basal ganglia in 30. Canavero S, Bonicalzi V, Paolotti R, Castellano G, Greco- Crasto S, Rizzo L, et al. Therapeutic extradural cortical Parkinson’s disease. Trends Neurosci 2000; 23(Suppl): S8– S19. stimulation for movement disorders: a review. 44. Dorval AD, Russo GS, Hashimoto T, Xu W, et al. Deep Neurological Research 2003; 25: 118–22. brain stimulation reduces neuronal entropy in the MPTP 31. Plaha P, Khan S, Gill SS. Bilateral stimulation of the caudal zona incerta nucleus for tremor control. J of Neurol primate model of Parkinson’s disease. J Neurophysiol 2008; 100(5): 2807–18. Neurosurg and Psychi 2008; 79: 504-513. 45. Lafreniere-Roula M, Darbin O, Hutchison WD, Wichmann 32. Anderson ME, Postupna N, Ruffo M. Effects of high- frequency stimulation in the internal globus pallidus on the activity of thalamic neurons in the awake monkey. J of Neurophys 2003; 89: 1150–60. 33. Montgomery Jr. EB. Effects of globus pallidus interna stimulation on human thalamic neuronal activity. Clinical Neurophys 2006; 117: 2691–2702. T, et al. Apomorphine reduces subthalamic neuronal entropy in parkinsonian patients. Exp Neurol 2010; 225: 455–58. 46. Bercovici H, Constantin P, Foias C, Manley OP. Exponential decay of the power spectrum of turbulence. J Stat Phys 1995; 80:579–602. 47. Doering CR, Titi ES. Exponential decay rate of the power spectrum or solutions of the Navier–Stokes equations. Phys Fluids 1995; 76:1384–90. 7 48. Aurell E, Boffetta G, Crisanti A, Paladin G, Vulpiani A. Growth of noninfinitesimal perturbations in turbulence. PRL 1996;77(7):1262–5. 49. Lapeyre G. Characterization of finite-time Lyapunov exponents and vectors in twodimensional turbulence. Chaos 2002;12(3):688–98. 50. Lin DC, Hughson RL. Modeling Heart Rate Variability in Healthy Humans: A Turbulence Analogy. Phys Rev Lett 2001; 86: 1650–53. 51. Amari S. Dynamics of pattern formation in lateral inhibition type neural fields. Biol Cyb 1977; 27: 77–87. 52. Amari S. Homogeneous nets of neuron-like elements. Biol Cyb 1975; 17: 211–20. 53. Ermentrout GB, Kleinfeld D. Traveling electrical waves in cortex: Insights from phase dynamics and speculation on a computational role. Neuron 2001; 29: 33–44. 54. Nunez PL. The brain wave equation: a model for the EEG. Math Biosci 1974; 21: 279–97. 55. Pinto DJ, ErmentroutGB. Spatially structured activity in synaptically coupled neuronal networks: I. Travelling fronts and pulses. SIAM J on Appl Math2001; 62: 206–25. 56. Gatev P, Darbin O, Wichmann T. Oscillations in the basal ganglia under normal conditions and in movement disorders, Movement Disord 2006; 21(10): 1566–77. 57. Tseng KY, Kasanetz F, Kargieman L, Riquelme LA, Murer MG. Cortical slow oscillatory activity is reflected in the membrane potential and spike trains of striatal neurons in rats with chronic nigrostriatal lesions, J Neurosci 2001; 21(16): 6430–39. 58. Stanford IM. Independent neuronal oscillators of the rat Globus Pallidus. J Neurophysiol 2003; 89: 1713–17. 59. Levy R, Dostrovsky JO, Lang AE, Sime E, Hutchison WD, Lozano AM. Effects of Apomorphine on subthalamic nucleus and Globus Pallidus Internus neurons in patients with Parkinson’s disease. J Neurophysiol 2001; 86: 249– 60. 60. Brown P. Bad oscillations in Parkinson’s disease. J Neural Transm 2006; 70(Suppl.): 27–30. 61. Baddeley R, Abbott LF, Booth MCA, Sengpiel F, Freeman T, Wakeman EA, et al. Responses of neurons in primary and inferior temporal visual cortices to natural scenes. Proc R Soc Lond B 1997; 264: 1775–83. 62. Bair W, Koch C, NewsomeW, Britten K. Power spectrum analysis of bursting cells in area MT in the behaving monkey. J Neurosci 1994; 74(5): 2870–92. 8
1906.00511
2
1906
2019-06-07T03:13:24
Deep learning from wristband sensor data: towards wearable, non-invasive seizure forecasting
[ "q-bio.NC", "eess.SP" ]
Seizure forecasting may provide patients with timely warnings to adapt their daily activities and help clinicians deliver more objective, personalized treatments. While recent work has convincingly demonstrated that seizure risk assessment is possible, these early approaches relied largely on complex, often invasive setups including intracranial electrocorticography, implanted devices and multi-channel EEG, which limits translation of these methods to broad clinical application. To facilitate broader adaptation of seizure forecasting in clinical practice, non-invasive, easily applicable techniques that reliably assess seizure risk, in combination with clinical information, are crucial. Wristbands that continuously record physiological parameters, including electrodermal activity, body temperature, blood volume pressure and actigraphy, may afford monitoring of autonomous nervous system function and movement relevant for such a task, hence minimizing potential complications associated with invasive monitoring, and avoiding stigma associated with bulky external monitoring devices on the head. Here, we use deep learning to analyze long-term, multi-modal wristband sensor data from 50 patients with epilepsy (total duration $>$1400 hours) to assess its capability to distinguish preictal from interictal states. Prediction performance is assessed using area under the receiver operating charateristic (AUC) and improvement over chance (IoC) based on F1 scores. Using one- and two-dimensional convolutional neural networks, we identified better-than-chance predictability in out-of-sample test data in 60\% of the patients in leave-one-out and 43\% of patients in pseudo-prospective approaches. These results provide a step towards developing easier to apply, non-invasive methods for seizure risk assessments in patients with epilepsy.
q-bio.NC
q-bio
Deep learning from wristband sensor data: towards wearable, non-invasive seizure forecasting Christian Meisel1,2∗ Rima El Atrache 1 Michele Jackson 1 Sarah Schubach 1 Claire Ufongene 1 Tobias Loddenkemper1 1 Boston Children's Hospital, Boston, USA 2 Department of Neurology, University Clinic Carl Gustav Carus, Dresden, Germany ∗ corresponding author email address: [email protected] 9 1 0 2 n u J 7 ] . C N o i b - q [ 2 v 1 1 5 0 0 . 6 0 9 1 : v i X r a 1 Seizure forecasting may provide patients with timely warnings to adapt their daily activities and help clinicians deliver more objective, personalized treatments. While recent work has convincingly demonstrated that seizure risk assessment is possi- ble, these early approaches relied largely on complex, often invasive setups includ- ing intracranial electrocorticography, implanted devices and multi-channel EEG, which limits translation of these methods to broad clinical application. To facilitate broader adaptation of seizure forecasting in clinical practice, non-invasive, easily applicable techniques that reliably assess seizure risk, in combination with clinical information, are crucial. Wristbands that continuously record physiological param- eters, including electrodermal activity, body temperature, blood volume pressure and actigraphy, may afford monitoring of autonomous nervous system function and movement relevant for such a task, hence minimizing potential complications associated with invasive monitoring, and avoiding stigma associated with bulky external monitoring devices on the head. Here, we use deep learning to analyze long-term, multi-modal wristband sensor data from 50 patients with epilepsy (to- tal duration >1400 hours) to assess its capability to distinguish pre- from interictal states. Prediction performance is assessed using area under the receiver operating charateristic (AUC) and improvement over chance (IoC) based on F1 scores. Using one- and two-dimensional convolutional neural networks, we identified better-than- chance predictability in out-of-sample test data in 60% of the patients in leave-one- out and 43% of patients in pseudo-prospective approaches. These results provide a step towards developing easier to apply, non-invasive methods for seizure risk assessments in patients with epilepsy. 2 Introduction Reliable methods to assess seizure risk could alleviate a major burden for epilepsy patients by providing timely warning or relief when seizure risk is high or low. From a clinician perspec- tive, robust seizure risk assessments are desirable because of their ability to improve treatment by optimizing dosing and timing of antiepileptic drug regimen by objective, personalized stan- dards, as well as by potentially enabling timely interventions to avert impending seizures.1 Following initial attempts,2 there has been a recent surge of studies demonstrating the pos- sibility of accurate seizure forecasting.3 -- 6 To this end, most studies have utilized either elec- trocorticography (ECoG) or scalp electroencephalography (EEG) as well as, to a lesser extent, electrocardiography (ECG), and have demonstrated that robust differentiation between preictal and interictal states is possible with a performance better than chance.7 -- 10 In order to make seizure risk assessments available for broader clinical use, however, methods that build on non-invasive, easily recordable data streams are desirable.11 Peripheral signals recorded us- ing wearable devices, such as wristbands, are particularly interesting in this respect since these signals permit continuous, non-invasive recording of several physiological parameters. At the same time, the compact design may limit the risk of stigmatization, are easy to apply, and may altogether increase patient adherence relevant for long-term ambulatory use. Continuous and simultaneous monitoring of a range of physiological parameters, such as electrodermal activity, body temperature, blood volume pressure and actigraphy, using wrist- bands is becoming increasingly available and permits close monitoring of autonomous nervous system function and movement,12 and may assist in the detection of generalized tonic-clonic seizures.13 Similar autonomous system measures may also provide information on detection of preictal patterns or states. Deep learning has been shown to exhibit strong classification performance 3 from complex feature sets.14 It therefore constitutes a promising technique to differentiate pre- from interictal states from complex, multi-modal wristband data. While more traditional ma- chine learning approaches rely on hand-designed feature sets, deep learning uses multiple layers of connections to perform classification tasks without the need of feature designing, which may be an advantage in relatively under-explored, multi-modal datasets, such as data from wristworn devices. In this work, we use convolutional neural networks (CNNs) on a unique dataset comprised of multi-modal wristband sensor data recorded from patients with epilepsy during multi-day in- hospital monitoring. Our aim is to evaluate the utility of this novel data and methods approach in its ability to differentiate pre- from interictal states as a prerequisite for future, non-invasive seizure forecasting methods. Materials and Methods Data recording and preprocessing We recruited patients with epilepsy admitted to the long-term video-EEG monitoring (LTM) unit and placed a biosensor wristband (E4, Empatica12) on either left or right wrist or ankle for long-term recording. For the purpose of this study we considered all patients with wristband recordings from 02/2015 until 11/2018. For the purpose of this study, we considered data from one wristband per patient only. When patient recording involved multiple wristbands (e.g. from wrist and ankle), we selected the data from the biosensor wristband with the longest total recording time for further analysis. A prerequisite for seizure risk assessment is the reliable distinction between pre- and inter- ictal states. We aimed to differentiate pre- from interictal states based on 30-second wristband recordings composed of six sensor data streams (electrodermal activity (EDA), accelerometer data in three dimensions, blood volume pressure (BVP), and temperature (TEMP); Fig. 1 A). 4 We considered a 30-second data segment as preictal if it occurred between 61 minutes and one minute prior to a seizure, thus leaving a one minute buffer prior to seizure onset (Fig. 1 B, red). Electrographic seizure onset was determined using video and EEG recordings. We analyzed all epileptic seizure types occurring in a patient, which included subclinical seizures, focal, primary and secondary tonic-clonic, myoclonic, clonic, tonic, atonic seizures and epileptic spasms (Fig- ure 4). Only seizures that occurred two hours or more from a preceding seizure were included to limit our analysis on lead seizures. 30-second data segments were classified as interictal if they occurred two hours or more from any seizure (Fig. 1 B, green). To allow stable recording conditions, we removed data from the first and last hour of each recording. Separation in training, validation and test data To evaluate classification performance on out of-sample test data, we used two separate ap- proaches, a leave-one-out approach and a pseudo-prospective approach. Leave-one-out approach A leave-one-out cross-validation approach has the advantage that all seizures can be used for training and validation. Here, we used a leave-one-out approach for each seizure. Similar to,15 if a subject had N seizures, (N−1) seizures were used for training/validation (75%/25% split in temporal order), and the remaining seizure was used for testing. This was done N times, so that all seizures were used for testing once. Interictal segments were randomly split into N parts, where (N − 1) parts were used for training/validation (75%/25% split in temporal order) and the remaining part was used for testing. Fig. 1 C (top) illustrates the leave-one-out separation using the first preictal (seizure) segment. The leave-one-out approach consequently required patients with at least two seizures which reduced the analysis to 50 patients (Figure 4). 5 Pseudo-prospective approach For seizure prediction under real-world conditions, the risk for seizures at a given time needs to be assessed using algorithms trained and validated on past data. A pseudo-prospective evalua- tion method that only evaluates an algorithm using out-of-sample data recorded at a later point in time than the data used for algorithm training and validation matches these conditions most closely. A pseudo-prospective evaluation setup benefits from large data sets with sufficient data for training and pseudo-prospective validation and testing. For this purpose, we separated data into non-overlapping, consecutive training, validation and test data, where validation and test data contained preictal segments from one seizure each and training data the remaining, preced- ing preictal and interictal periods (Fig. 1 C, bottom). The exact cut-off between validation and test data was chosen so that both contained an equal amount of interictal data. This procedure guaranteed that training and validation data always chronologically preceded test data and that training, validation and test data were chronologically separated. This constraint on data, which required patients to have had at least three seizures, reduced the analysis to 21 patients whose data could be separated for such a pseudo-prospective approach. Neural networks and training We primarily used one-dimensional convolutional neural networks (CNNs), as they have been shown to provide robust classification performance based on multi-dimensional timeseries data.14 To use the wristband sensor data in a one-dimensional CNN (1DConv), data was down-sampled to 4 Hz for all sensors in order to provide the same vector length for each 30-second segment. Figure 5 shows a summary of the CNN parameters used. We also compared our results to a two-dimensional convolutional neural network (2DConv) that used the power spectral density (PSD, FFT routine; 30-second, non-overlapping Hanning windows) of signals sampled at 64 Hz (if originally sampled below 64 Hz, signals were upsam- 6 pled by repeating values). PSD data were then reshaped in 31 by 31 images of depth 6 for use in neural networks. The use of 2DConv networks using PSD was motivated by encouraging results of similar approaches using EEG and ECoG data.15, 16 We used balanced learning to handle imbalanced training sets, repeating either pre- or in- terictal segments in the training set until a 50-50 preictal-interictal balance was achieved. All networks were trained for 1000 epochs. To limit CNNs from overfitting, we kept the CNN architecture simple and shallow. Performance metrics and statistical tests Performance was assessed on out-of-sample test data using the area under the receiver operating curve (AUC) and improvement over chance based on F1 scores (IoC). AUC is commonly used to assess performance in classification problems. AUC scores were obtained for ten network executions (pseudo-prospective approach) or all folds (leave-one-out approach, five network executions each). A one-sample t-test was used to compare AUC-values against chance (i.e., AUC=0.5). The F1 score is the harmonic mean between precision and recall and, as a single metric, evaluates how well a classification algorithm performs on the minority class. The use of F1 scores as an additional method was motivated by the observation that F1 scores may provide more reliable performance estimates in skewed, imbalanced data sets,17, 18 such as in classifica- tion problems in epilepsy where interictal data typically outnumber preictal data. F 1T est scores were obtained for ten network executions (pseudo-prospective approach) or all folds (leave-one- out approach, five network executions each). To compare our algorithms to a chance predictor, the F1 score was obtained for random predictions (i.e., randomly shuffled test set lablels, which maintains the numbers of pre- and interictal classifications but unlinks any correlation to the data) averaged across 1000 randomizations (pseudo-prospective approach; 1000 randomization 7 per fold for the leave-one-out approach) resulting in a mean F 1Chance score. Improvement over chance (IoC) was then defined as IoC = F 1T est−F 1Chance metric in this article. A one-sample t-test was used to compare IoC-values against chance (i.e., and used as the main performance 1−F 1Chance IoC=0). IoC has been repeatedly used in epilepsy forecasting research as a meaningful metric to characterize algorithm performance.2, 7, 16 Results We assessed the utility of CNNs in distinguishing pre- from interictal states using long-term, multi-modal wristband sensor data12 obtained during epilepsy monitoring. Neural network based assessments were done for each patient individually and for two approaches, a leave- one-out cross-validation approach and a pseudo-prospective approach (Fig. 1). Leave-one-out approach The Leave-one-out approach required patient datasets to contain at least two seizures. With this constraint, we analyzed a total of 50 patients with a total of 157 seizures and a recording time of 1425.92 hours (1270.18 hours interictal, 155.74 hours preictal; Figure 4). Each fold of the leave-one-out approach was executed five times, and average results with standard deviations were reported for each patient. Area under the receiver operating curve (AUC) values and improvement over chance (IoC) values using the 1DConv network are de- picted along with seizure number and data duration for all patients in figure 2. Overall, a prediction better than chance was obtained for more than half of the patients (Fig. 2 A, B, blue markers; AUC significant for 30/50 patients, AU C = 0.714 ± 0.121; IoC significant for 29/50 patients, IoC = 0.377 ± 0.229). In comparison, the 2DConv network did not exhibit an overall better performance (AUC significant for 26/50 patients, AU C = 0.595± 0.069; IoC significant for 29/50 patients, IoC = 0.164 ± 0.131). 8 Pseudo-prospective approach Although our data here is comparably short, we attempted a pseudo-prospective evaluation to assess our approach under more realistic conditions. Requiring validation and test data to be chronologically following training data with at least one seizure each reduced our analysis to 21 patients. Networks were executed ten times, and average results with standard deviations were reported (Fig. 3 A, B, blue markers; AUC significant for 9/21 patients, AU C = 0.627 ± 0.090; IoC significant for 8/21 patients, IoC = 0.221 ± 0.144). The 2DConv network again did not exhibit an overall better performance (AUC significant for 9/21 patients, AU C = 0.572±0.062; IoC significant for 13/21 patients, IoC = 0.094 ± 0.061). Discussion Here we assessed the utility of physiological sensor data recorded from a wristband to estimate seizure risk. The ability to robustly differentiate preictal from an interictal states is a neces- sary prerequisite for any seizure forecasting or seizure risk assessment.2 For this classification process, we used convolutional neural networks which have demonstrated outstanding perfor- mance in classification tasks in several domains,14 more recently also including epilepsy.15, 16 Our work is motivated by the potential benefits for patients and clinicians from a robust seizure gauge. Forecasting seizures would provide patients with timely warning to adapt daily activ- ities and allow clinicians to titrate therapies and develop novel interventions that potentially could prevent impending seizures.1, 19 Peripheral sensor data that can be recorded easily and non-invasively with a wristband would be desirable for such a purpose since approaches based on ECoG7 or a large number of scalp EEG channels11 limit broad clinical application. Of the patients included in our analysis, about half displayed a significant improvement over chance (IoC) for both evaluation schemes, leave-one-out and pseudo-prospective approaches. 9 On the one hand, these performance values may not appear as strong as what is reported in other recent studies where the majority of patients exhibited predictability levels better than chance.15, 16 On the other hand, these results suggest that seizure forecasting might also be fea- sible with relatively short, noisy, multi-modal signals recorded from wristbands far away from the brain. The better-than-chance classification performance in about half of the patients was obtained despite the comparably brief duration of data, where training sometimes only involved one or two seizures, and the variability in seizure types, data duration, age and wristband lo- cation. Apart from the criteria to label pre- and interictal data, we used no other preselection or preprocessing, but instead included the data in raw format "as is" in an attempt to maximize the transferability of our approach to real-world, noisy conditions. Predictability performance across patients did not depend on wristband location, overall duration of data, or seizure type. Seizure forecasting builds on the notion that a preictal state, during which a seizure is more likely to occur soon, can be reliably distinguished from interictal states. To this end, most stud- ies have focused on data recorded either from ECoG and EEG or from ECG. ECG has thus been a long-standing example that peri- and preictal changes can not only be detected within the cen- tral nervous system but are also reflected in a variety of cardiac effects.20 -- 22 Cardiac activity is controlled by parasympathetic and sympathetic branches of the autonomic nervous system, with the former producing an inhibitory response and the latter producing an excitatory response on heart rate.23 Preictal changes in brain activity that occur in or propagate to autonomic control centers may affect this autonomic balance and, consequently, affect cardiac activity during the leadup to a seizure. A recent study that compared the information content in ECoG, EEG and ECG in terms of identifying preictal states found that single-channel ECG contains a compara- ble amount of information to multi-channel EEG,24 which highlights the relevance of peripheral sensors for seizure forecasting. Autonomous nervous system changes are captured by the wrist- band sensors used in this study in several ways. Electrodermal activity is known to be sensitive 10 to sympathetic innervation. Blood volume pressure curves contain information about heart rate which is controlled by the parasympathetic and sympathetic interplay. Similarly, body temper- ature is known to be maintained by the autonomic nervous system. The approach proposed in this study builds on monitoring these autonomous nervous system functions along with actig- raphy, which indirectly also monitors resting periods and sleep, and therefore pioneers seizure forecasting capabilities based on such multi-modal sensor data, going beyond more traditional ECoG/EEG and ECG approaches. In our approach, we used the same model for all patients, albeit models were trained for each patient individually. While it is possible that model hyperparameters individualized for each patient might bring about better performances, we chose to have the same model architecture across patients that could potentially be implemented "out-of-the-box" in future prospective settings. Previous studies using CNNs for seizure forecasting have relied on power spectral densities converted to two-dimensional images used in neural networks.15, 16 In contrast, our approach relied on one-dimensional convolutional networks where we used raw sensor data for inputs directly, thus requiring less preprocessing. Although we did not do an extensive comparison in terms of model hyperparameters, we found that, at least for the two models chosen here, the 1DConv networks using raw data performed slightly better than the 2DConv network using the data's power spectrum. Our evaluation used two different approaches, each of which has certain advantages and disadvantages. Leave-one-out approaches have the advantage that all seizures can be used for training and validation to obtain an estimate of algorithm performance. This can be desirable particularly for relatively short data durations, such as ours. For this reason, leave-one-out cross- validation approaches have been used to assess seizure prediction and detection performance in epilepsy research.13, 15 We here applied a leave-one-out approach to preictal intervals belonging to different seizures in order to assess the algorithmic ability to differentiate pre- from interictal 11 states. For a forecasting algorithm to be used in real-world conditions to assess future seizure risk, classification of incoming data into pre- and interictal classes has to be performed in a prospective manner using data recorded after the training and validation periods. A pseudo- prospective assessment benefits particularly from large data sets with sufficient data for training and pseudo-prospective validation and testing.16 Results need to be interpreted in the setting of data acquisition. One limitation of our study is the relative short duration of recordings which covered only a few days of continuous data per patient. Training data benefits from long periods of data where algorithms can better learn to generalize and which gives a more realistic account of seizure forecasting capabilities. How- ever, the current dataset is unique in the sense that it contains multi-modal sensor data over several days from a relatively large number of epilepsy patients. The better-than-chance pre- dictability in about half of the patients in this study is therefore encouraging for future, longer trials using these sensors. Another limitation is the absence of benchmarks to compare our ap- proach to. While we compared our results to chance predictors in out-of-sample test data, the uniqueness and novelty of the current dataset limits more comprehensive comparison to other approaches. There is growing awareness of the benefits of creating data warehouse ecosystems that allow rigorous and continuous reevaluation and benchmarking by making data and algo- rithms available to many researchers.25 We expect that these open-science efforts will increase the reproducibility and help benchmark and improve algorithms, such as the ones proposed in the current study, in the future. Finally, other clinically meaningful metrics, such as false alarm rate, or time between seizure warning and actual seizure, that characterize an algorithm's fore- casting performance may be desirable. These metrics, however, build on top of classifications and require further post-processing. Due to the brevity of the data and the absence of other relevant benchmarks to compare our approach to, we did not consider this useful in the current context. Instead our main goal in this study was more basic, to assess whether the combination 12 of peripheral wristband sensor data and deep learning might be able to differentiate pre- from interictal states with a better-than-chance performance. For this purpose, AUC and IoC, which have both been used in several other forecasting studies,2, 7, 16, 25, 26 are valid metric which also allow future comparison and benchmarking. Seizure forecasting is likely to bring about notable benefits for epilepsy patients and clini- cians. In order to make seizure forecasting available for broad use, non-invasive, easily appli- cable techniques are greatly desirable. We here assessed the capability of multi-modal wrist- band sensor data in combination with deep learning to reliably distinguish pre- from interictal states. Our results demonstrate a better-than-random predictability in about half the patients, even when a pseudo-prospective approach on out-of-sample test data was taken. Future, more long-term studies should help to validate the utility of this approach. References 1 S. B. Dumanis, J. A. French, C. Bernard, G. A. Worrell, and B. E. Fureman. Seizure Forecast- ing from Idea to Reality. Outcomes of the My Seizure Gauge Epilepsy Innovation Institute Workshop. eNeuro, 4(6), 2017. 2 F. Mormann, R. G. Andrzejak, C. E. Elger, and K. Lehnertz. Seizure prediction: the long and winding road. Brain, 130:314 -- 333, 2007. 3 D. R. Freestone, P. J. Karoly, and M. J. Cook. A forward-looking review of seizure prediction. Curr. Opin. Neurol., 30(2):167 -- 173, 04 2017. 4 Melbourne University AES/MathWorks/NIH Seizure Prediction. https://www.kaggle.com/c/melbourne-university-seizure-prediction. Accessed: 2016-10-28. 13 5 F. Mormann and R. G. Andrzejak. Seizure prediction: making mileage on the long and winding road. Brain, 139(Pt 6):1625 -- 1627, Jun 2016. 6 L. Kuhlmann, K. Lehnertz, M. P. Richardson, B. Schelter, and H. P. Zaveri. Seizure prediction - ready for a new era. Nat Rev Neurol, 14(10):618 -- 630, Oct 2018. 7 M. J. Cook, T. J. O'Brien, S. F. Berkovic, M. Murphy, A. Morokoff, G. Fabinyi, W. D'Souza, R. Yerra, J. Archer, L. Litewka, S. Hosking, P. Lightfoot, V. Ruedebusch, W. D. Sheffield, D. Snyder, K. Leyde, and D. Himes. Prediction of seizure likelihood with a long-term, im- planted seizure advisory system in patients with drug-resistant epilepsy: a first-in-man study. Lancet Neurol, 12(6):563 -- 571, Jun 2013. 8 K. Lehnertz, H. Dickten, S. Porz, C. Helmstaedter, and C. E. Elger. Predictability of uncon- trollable multifocal seizures - towards new treatment options. Sci Rep, 6:24584, Apr 2016. 9 K. Fujiwara, M. Miyajima, T. Yamakawa, E. Abe, Y. Suzuki, Y. Sawada, M. Kano, T. Mae- hara, K. Ohta, T. Sasai-Sakuma, T. Sasano, M. Matsuura, and E. Matsushima. Epileptic Seizure Prediction Based on Multivariate Statistical Process Control of Heart Rate Variability Features. IEEE Trans Biomed Eng, 63(6):1321 -- 1332, 06 2016. 10 D. H. Kerem and A. B. Geva. Forecasting epilepsy from the heart rate signal. Med Biol Eng Comput, 43(2):230 -- 239, Mar 2005. 11 A. Schulze-Bonhage, F. Sales, K. Wagner, R. Teotonio, A. Carius, A. Schelle, and M. Ihle. Views of patients with epilepsy on seizure prediction devices. Epilepsy Behav, 18(4):388 -- 396, Aug 2010. 12 M. Z. Poh, T. Loddenkemper, N. C. Swenson, S. Goyal, J. R. Madsen, and R. W. Picard. Continuous monitoring of electrodermal activity during epileptic seizures using a wearable sensor. Conf Proc IEEE Eng Med Biol Soc, 2010:4415 -- 4418, 2010. 14 13 M. Z. Poh, T. Loddenkemper, C. Reinsberger, N. C. Swenson, S. Goyal, M. C. Sabtala, J. R. Madsen, and R. W. Picard. Convulsive seizure detection using a wrist-worn electrodermal activity and accelerometry biosensor. Epilepsia, 53(5):e93 -- 97, May 2012. 14 J. Schmidhuber. Deep learning in neural networks: an overview. Neural Netw, 61:85 -- 117, Jan 2015. 15 N. D. Truong, A. D. Nguyen, L. Kuhlmann, M. R. Bonyadi, J. Yang, S. Ippolito, and O. Kave- hei. Convolutional neural networks for seizure prediction using intracranial and scalp elec- troencephalogram. Neural Netw, 105:104 -- 111, Sep 2018. 16 I. Kiral-Kornek, S. Roy, E. Nurse, B. Mashford, P. Karoly, T. Carroll, D. Payne, S. Saha, S. Baldassano, T. O'Brien, D. Grayden, M. Cook, D. Freestone, and S. Harrer. Epilep- tic Seizure Prediction Using Big Data and Deep Learning: Toward a Mobile System. EBioMedicine, 27:103 -- 111, Jan 2018. 17 Nitesh V. Chawla. Data Mining for Imbalanced Datasets: An Overview, pages 853 -- 867. Springer US, Boston, MA, 2005. 18 L. A. Jeni, J. F. Cohn, and F. De La Torre. Facing Imbalanced Data Recommendations for the Use of Performance Metrics. Int Conf Affect Comput Intell Interact Workshops, 2013:245 -- 251, 2013. 19 M. O. Baud and V. R. Rao. Gauging seizure risk. Neurology, 91(21):967 -- 973, Nov 2018. 20 J. M. VAN BUREN. Some autonomic concomitants of ictal automatism; a study of temporal lobe attacks. Brain, 81(4):505 -- 528, Dec 1958. 21 D. W. Marshall, B. F. Westmoreland, and F. W. Sharbrough. Ictal tachycardia during temporal lobe seizures. Mayo Clin. Proc., 58(7):443 -- 446, Jul 1983. 15 22 P. E. Smith, S. J. Howell, L. Owen, and L. D. Blumhardt. Profiles of instant heart rate during partial seizures. Electroencephalogr Clin Neurophysiol, 72(3):207 -- 217, Mar 1989. 23 B. F. Robinson, S. E. Epstein, G. D. Beiser, and E. Braunwald. Control of heart rate by the autonomic nervous system. Studies in man on the interrelation between baroreceptor mecha- nisms and exercise. Circ. Res., 19(2):400 -- 411, Aug 1966. 24 C. Meisel and K. Bailey. Deep learning identifies signal-dependent information about the preictal state: a comprehensive assessment across ECoG, EEG and EKG. submitted, 2019. 25 B. H. Brinkmann, J. Wagenaar, D. Abbot, P. Adkins, S. C. Bosshard, M. Chen, Q. M. Tieng, J. He, F. J. Munoz-Almaraz, P. Botella-Rocamora, J. Pardo, F. Zamora-Martinez, M. Hills, W. Wu, I. Korshunova, W. Cukierski, C. Vite, E. E. Patterson, B. Litt, and G. A. Worrell. Crowdsourcing reproducible seizure forecasting in human and canine epilepsy. Brain, 139(Pt 6):1713 -- 1722, 06 2016. 26 L. Kuhlmann, P. Karoly, D. R. Freestone, B. H. Brinkmann, A. Temko, A. Barachant, F. Li, G. Titericz, B. W. Lang, D. Lavery, K. Roman, D. Broadhead, S. Dobson, G. Jones, Q. Tang, I. Ivanenko, O. Panichev, T. Proix, M. Nahlik, D. B. Grunberg, C. Reuben, G. Worrell, B. Litt, D. T. J. Liley, D. B. Grayden, and M. J. Cook. Epilepsyecosystem.org: crowd-sourcing reproducible seizure prediction with long-term human intracranial EEG. Brain, 141(9):2619 -- 2630, Sep 2018. Acknowledgements This work was supported in part by the Epilepsy Research Fund (ERF). CM acknowledges sup- port by a NARSAD Young Investigator Grant from the Brain & Behavior Research Foundation. 16 Ethical Publication Statement We confirm that we have read the Journals position on issues involved in ethical publication and affirm that this report is consistent with those guidelines. Disclosure None of the authors has any conflict of interest to disclose. TL received unrelated research support and device donations from Empatica, Inc. 17 Figure 1: Schematic outline of data processing in order to predict pre- and interictal states. A, continuous recording of timeseries data. B, data separation into preictal and interictal segments. C, separation into training, validation and test data for each approach, leave-one-out (example shown for first seizure) and pseudo-prospective. D, training, validation and out-of-sample test- ing using convolutional neural networks. 18 AUTOCORRpredict seizures with theory + deep learningCRITICAL SLOWVARIANCEextract predictive EEG biomarkers no seizureseizure soonBVP blood volume pressureAUTOCORRpredict seizures with theory + deep learningCRITICAL SLOWVARIANCEextract predictive EEG biomarkers no seizureseizure soonDeep LearningTimeseries DataEDA electrodermal activityTEMP temperatureACC-X accelerometer in x-dimensionACC-Y accelerometer in y-dimensionACC-Z accelerometer in z-dimensionValidateTrainAUTOCORRpredict seizures with theory + deep learningCRITICAL SLOWVARIANCEextract predictive EEG biomarkers no seizureseizure soonLeave-One-Out ApproachTime [Hours]240PreictalInterictalData Separation In Pre-/InterictalTime [Hours]240TrainVal.PreictalInterictalPseudo-Prospective ApproachTestTime [Hours]240TestAUTOCORRpredict seizures with theory + deep learningCRITICAL SLOWVARIANCEextract predictive EEG biomarkers no seizureseizure soonAUTOCORRpredict seizures with theory + deep learningCRITICAL SLOWVARIANCEextract predictive EEG biomarkers no seizureseizure soonDeep LearningAUTOCORRpredict seizures with theory + deep learningCRITICAL SLOWVARIANCEextract predictive EEG biomarkers no seizureseizure soonABCD Figure 2: Prediction performance for the leave-one-out approach using a one-dimensional convolutional network (1DConv, n=50 patients). A, area under the receiver operating curve (AUC; blue markers indicate a performance significantly better than chance). B, improvement over chance (IoC). C, seizure count. D, total duration of data analyzed. 19 IoC (Leave-One-Out Approach)Patient IDChanceMaxNumber Of Seizures AnalysedHours AnalysedBCDAAUC (Leave-One-Out Approach)p≤0.01p≤0.05* Figure 3: Prediction performance for the pseudo-prospective approach using a one-dimensional convolutional network (1DConv, n=21 patients). A, area under the receiver operating curve (AUC; blue markers indicate a performance significantly better than chance). B, improvement over chance (IoC). C, seizure count. D, total duration of data analyzed. 20 IoC (Pseudo-Prospective Approach)Patient IDChanceMaxNumber Of Seizures AnalysedHours AnalysedBCDp≤0.01p≤0.05*AAUC (Pseudo-Prospective Approach) Figure 4: Summary of patient characteristics. 21 Patient dataPatient ID101413130.812138151311542291581313121522103388 390 392 399 410 411 417 418 419 421423 425 427 428 432 440 443 444 446 466 467 475 482 493 496 506Age [Years]LocationWristband1483211653671090.669270.4421479111911117No. of SeizuresPreictal[Minutes]Interictal[Minutes]180.0 180.0 120.0 240.0 120.0 120.0 275.5 180.0 180.0 180.0 120.0 120.0 120.0 240.0 120.0 300.0 181.5 64.5 120.0 240.0 180.0 180.0 120.0 240.0 180.0 480.0597.5 914.5 2243.5 2356.0 204.0 2228.5 2714.0 545.5 1983.5 1753.5 2270.5 807.0 1370.0 1427.5 185.5 1113.5 1764.5 2255.5 4536.5 2986.5 1890.0 2116.5 3806.5 3215.0 187.0 832.0right wristleft wrist right ankle left wrist right ankle right wrist left ankle right ankle right wrist left wrist left wrist left wrist left ankle left ankle right ankle left ankle right ankle right wrist left wrist right wrist right wrist left ankle left ankle left ankle right wrist right wristPatient IDAge [Years]LocationWristbandNo. of SeizuresPreictal[Minutes]Interictal[Minutes]192 198 210 212 221 224 236 243 263 284 296 302 303 309 329 330353 356 365 370 372 378 380 3873 3 2 4 2 2 5 3 3 3 2 2 2 4 2 5 4 2 2 4 3 3 2 4 3 8103 2 3 4 2 2 3 2 2 6 4 5 2 2 2 2 4 3 2 2 2 2 3657.0138.5 120.0 180.0 214.5 120.0 120.0 180.0 120.0 120.0 360.0 240.0 300.0 119.5 120.0 120.0 120.0 193.0 180.0 120.0 120.0 120.0 120.0 180.0 3398.5 1136.0 864.5 395.0 1160.5 2645.5 98.0 570.5 391.0 753.0 1134.5 1478.0 2541.0 617.0231.5 16.5 1660.5 526.5 1616.5 702.5 474.0 22.5 3009.0 4464.0 right wrist left ankle right ankle right ankle left wristleft wristleft ankle left wrist left wrist left ankle right ankle right ankle left ankle right wrist right ankle left ankle left ankle left ankle right ankle right wrist right ankle left ankle right wrist right wrist Figure 5: One-dimensional convolutional (1DConv) network topology used. Learning rate was set to 0.0001. 22 Network architectureLayer numberLayer typeFilters / Kernel size or NodesLayer parametersConv1DDropoutConv1DMaxPool1DConv1DDropoutConv1DAvgPool1DDropoutDense1234567891050/10NA50/10380/102080/10NANA1activation = ReLUdropout rate = 0.7activation = ReLUNAactivation = ReLUdropout rate = 0.7activation = ReLUNAdropout rate = 0.7activation = sigmoid
1805.03601
3
1805
2018-11-30T19:57:40
Finite size effects for spiking neural networks with spatially dependent coupling
[ "q-bio.NC" ]
We study finite-size fluctuations in a network of spiking deterministic neurons coupled with non-uniform synaptic coupling. We generalize a previously developed theory of finite size effects for uniform globally coupled neurons. In the uniform case, mean field theory is well defined by averaging over the network as the number of neurons in the network goes to infinity. However, for nonuniform coupling it is no longer possible to average over the entire network if we are interested in fluctuations at a particular location within the network. We show that if the coupling function approaches a continuous function in the infinite system size limit then an average over a local neighborhood can be defined such that mean field theory is well defined for a spatially dependent field. We then derive a perturbation expansion in the inverse system size around the mean field limit for the covariance of the input to a neuron (synaptic drive) and firing rate fluctuations due to dynamical deterministic finite-size effects.
q-bio.NC
q-bio
Finite size effects for spiking neural networks with spatially dependent coupling Si-Wei Qiu1 and Carson C. Chow1 1Mathematical Biology Section, LBM, NIDDK, NIH (Dated: December 4, 2018) 8 1 0 2 v o N 0 3 ] . C N o i b - q [ 3 v 1 0 6 3 0 . 5 0 8 1 : v i X r a 1 Abstract We study finite-size fluctuations in a network of spiking deterministic neurons coupled with non- uniform synaptic coupling. We generalize a previously developed theory of finite size effects for globally coupled neurons with a uniform coupling function. In the uniform coupling case, mean field theory is well defined by averaging over the network as the number of neurons in the network goes to infinity. However, for nonuniform coupling it is no longer possible to average over the entire network if we are interested in fluctuations at a particular location within the network. We show that if the coupling function approaches a continuous function in the infinite system size limit then an average over a local neighborhood can be defined such that mean field theory is well defined for a spatially dependent field. We then use a path integral formalism to derive a perturbation expansion in the inverse system size around the mean field limit for the covariance of the input to a neuron (synaptic drive) and firing rate fluctuations due to dynamical deterministic finite-size effects. I. INTRODUCTION The dynamics of neural networks have traditionally been studied in the limit of very large numbers of neurons, where mean field theory can be applied, e.g. [1, 2, 4, 11, 16, 19, 20, 25, 26, 30] or for a small number of neurons, where traditional dynamical systems approaches can be used, e.g. [15, 22, 24]. The intermediate regime of large but finite numbers of neurons can have interesting properties that are independent of the small and infinite system limits [6, 10, 12, 14, 17, 18, 23, 29]. However, these previous works have not fully explored fluctuations due to finite-size effects at specific locations within the network when all the neurons receive nonhomogeneous input from other neurons because of nonuniform coupling. Here, we consider finite-size effects in a network of spiking neurons with nonuniform synaptic coupling. Previously [6, 10, 18], a perturbation expansion in the inverse network neuron number had been developed for networks with global spatially uniform coupling and we generalize that theory to include nonuniform coupling. We first show that mean field theory in the infinite nonuniform system limit can be realized in a single network if a spatial metric can be imposed on the network and the coupling function is a continuous function of this distance measure. We then analyze finite-size fluctuations around such mean field solutions using a path integral formalism to derive a perturbation expansion in the inverse network 2 neuron number for the spatially dependent covariance function for the synaptic drive and spatially dependent neuron firing rate. II. COUPLED NEURON MODEL Consider a network of N theta neurons (phase reduction of quadratic integrate-and-fire neurons [15]) on a one dimensional periodic domain of size L although the theory can be applied to any domain. The network obeys the following deterministic microscopic equations: θi = 1 − cos θi + (Ii + ui(t))(1 + cos θi) N(cid:88) j=1 ui = L N wijsj (cid:88) sj = −βsj + β δ(t − tl j) (2.1) (2.2) (2.3) l where θi is the phase of neuron i, ui is the synaptic drive to neuron i, Ii is the external input to neuron i, β is the decay constant of the synaptic drive, sj is the time dependent synaptic input from neuron j, and tl j represents the spike times when the phase of neuron j crosses π. sj rises instantaneously when neuron j spikes and relaxes to zero with a time constant of 1/β. The synaptic drive represents the total time dependent synaptic input where the contribution from each neuron is weighted by the synaptic coupling function wij (a real N × N matrix). When Ii + ui > 0, the neuron receives suprathreshold input and θi will progress in time. When it passes π, the neuron is said to spike. When Ii + ui < 0 the neuron receives subthreshold input and the phase will approach a fixed point. The theta neuron is the normal form of a Type I spiking neuron near the bifurcation point to firing [15]. By linearity, the synaptic drive obeys the more convenient form of ui = −βui + β L N We define an empirical density N(cid:88) (cid:88) δ(t − tl j) wij j=1 l ηj(θ, t) = δ(θ − θj(t)) (2.4) (2.5) the sum of a spike train as(cid:80) that assigns a point mass to the phase of each neuron in the network. Hence, we can write θjθj =π = 2 j) = ηj(π, t) θjθj =π. For the theta model, l δ(t − tl 3 and thus we can then rewrite (2.4) as ui = −βui + 2β N(cid:88) j=1 L N wijηj(π, t) (2.6) Neuron number is conserved so the neuron density formally obeys a conservation (Klimon- tovich) equation [10]: ∂tηi(θ, t) + ∂θFi(θ, ui)ηi(θ, t) = 0 (2.7) where Fi(θ, ui) = 1 − cos θ + (1 + cos θ)(Ii + ui). The Klimontovich equation together with (2.6) fully describes the system. However, it is only a formal definition since η is not in general differentiable. In the following, we develop a method to regularize the Klimontovich equation so that desired quantities can be calculated. III. MEAN FIELD THEORY The Klimontovich equation (2.7) only exists in a weak sense. We can regularize it by taking a suitable average over an ensemble of initial conditions: ∂t(cid:104)ηi(θ, t)(cid:105) + ∂θ(cid:104)Fi(θ, ui)ηi(θ, t)(cid:105) = 0 (3.1) This equation is not closed because it involves covariances such as (cid:104)ηη(cid:105), which in turn depend on higher order cumulants in a BBGKY hierarchy [6, 10, 18]. This hierarchy can be rendered tractable if we can truncate it. Mean field theory truncates the hierarchy at first order by assuming that all cumulants beyond the first are zero so we can write ∂tρi(θ, t) + ∂θFi(θ, ai)ρi(θ, t) = 0 (3.2) where ai = (cid:104)ui(cid:105) and ρi = (cid:104)ηi(cid:105). The full set of closed mean field equations are given by ∂tρi(θ, t) + ∂θ [1 − cos θ + (Ii + ai)(1 + cos θ)] ρi(θ, t) = 0 ai = −βai + 2β L N wijρj(π, t) (3.3) (cid:88) j Although we can always write down the mean field equations (3.3), it is not clear that a given network would obey it in the infinite N limit. In previous work [9, 10, 25], it was shown that mean field theory applies to a network of coupled oscillators with uniform coupling in 4 the infinite N limit. However, it is not known when or if mean field theory applies for nonuniform coupling. To see this, consider first the stationary system ∂θ [1 − cos θ + (Ii + ui)(1 + cos θ)] ηi(θ) = 0 (cid:88) j ui = 2 L N wijηj(π) (3.4) (3.5) with uniform coupling, wij = w, and uniform external input, Ii = I. If the neurons are initialized with random phases and remain asynchronous then we can suppose that in the limit of N → ∞ the quantity (cid:88) j ρ(π) = L N ηj(π) (3.6) converges to an invariant quantity [9, 10, 25]. This then implies that ui = 2wρ ≡ a is also a constant. Thus each neuron will have identical inputs so if we apply the network averaging (cid:80)N operator L N i=1 to (3.4) we obtain ∂θ [1 − cos θ + (I + a)(1 + cos θ)] ρ(θ) = 0 a = 2wρ(π) (3.7) (3.8) Covariances vanish and mean field theory is realized in the infinite network limit. Given that the drive equation (2.6) is linear, the time dependent mean field theory will similarly hold in the large N limit. In the case where wij is not uniform, covariances are not guaranteed to vanish and an infinite network need not obey mean field theory. Our goal is to find conditions such that mean field theory applies. Again, consider the stationary equations (3.4) and (3.5). Now, instead of averaging over the entire domain, take a local interval around j, [j − cN/2, j + cN/2], where c < 1 is a constant that can depend on N and we map j − cN/2 < 1 to N + j − cN/2 and j + cN/2 > N to j + cN/2 − N . We want to express our mean field equation in terms of the locally averaged empirical density 2(cid:88) j+ cN k=j− cN 2 ρj = 1 cN ηk (3.9) If cN → ∞ for N → ∞ then it is feasible that the local empirical density can be invariant (to random initial conditions) and correlations can vanish; we seek conditions on the coupling 5 for which this is true. Inserting (3.5) into (3.4), and taking the local average yields (cid:34) 2(cid:88) k+ cN i=k− cN 2 1 cN 1 − cos θ + ∂θ (cid:32) (cid:33) L N Ii + 2 N(cid:88) j=1(cN )−1(cid:80)j+ cN j=1 wijηj(π) = (cid:80)N wijηj(π) j=1 2 Consider the identity (cid:80)N (cid:35) We immediately see that correlations can arise from the sums over the product of ηj(π)ηi(θ). (1 + cos θ) ηi(θ) = 0 (3.10) periodic boundary conditions. For nonperiodic boundary conditions there will be an edge contribution but this should be negligible in the large network limit. Using this summation l=j− cN 2 wilηl(π), which is exact for identity, we can rewrite the sum as N(cid:88) 1 cN 2(cid:88) k+ cN i=k− cN 2 wijηj(π)ηi(θ) = Rjk = 1 (cN )2 + 1 (cN )2 j+ cN 2(cid:88) 2(cid:88) 2 j+ cN l=j− cN k+ cN 2(cid:88) 2(cid:88) i=k− cN 2 k+ cN l=j− cN 2 i=k− cN 2 N(cid:88) j=1 (wkjρj(π)ρk(θ) + Rjk) (3.11) (wij − wkj)ρj(π)ηi(θ) (wil − wij)ηl(π)ηi(θ) (3.12) j=1 where the remainder carries the correlations. Mean field theory is valid in the N → ∞ limit if Rjk vanishes. Its magnitude obeys (wij − wkj)ρj(π)ηi(θ) (wil − wij)ηl(π)ηi(θ) (cid:12)(cid:12)(cid:12)(cid:12) + (cid:12)(cid:12)(cid:12)(cid:12) 1 (cN )2 2(cid:88) j+ cN 2(cid:88) k+ cN l=j− cN 2 i=k− cN 2 (cid:12)(cid:12)(cid:12)(cid:12) Rjk ≤ (cN )2 (cid:12)(cid:12)(cid:12)(cid:12) + i=k− cN 2 k+ cN 1 j+ cN l=j− cN 2(cid:88) (cid:18) 1 (cid:18) 1 2(cid:88) 2(cid:88) 2(cid:88) 2 j+ cN 2 k+ cN i=k− cN cN (cid:19) 2(cid:88) k+ cN ηi(θ) (cN )2 l=j− cN 2 i=k− cN 2 ≤ρj(π) since the density is nonnegative. Applying (3.9) then leads to (cid:32) i∈(k− cN sup 2 ,k+ cN 2 ) (cid:19) (ηl(π)ηi(θ)) wij − wkj i∈(k− cN 2 ,k+ cN sup 2 ),l∈(j− cN 2 ,j+ cN 2 ) wil − wij (3.13) (cid:33) wil − wij (3.14) R ≤ρj(π)ρk(θ) i∈(k− cN sup 2 ,k+ cN 2 ) i∈(k− cN 2 ,k+ cN sup 2 ),l∈(j− cN 2 ,j+ cN 2 ) wij − wkj + 6 We introduce a distance measure z = iL/N , z(cid:48) = jL/N ,z(cid:48)(cid:48) = kL/N , z(cid:48)(cid:48)(cid:48) = lL/N and write ρi(θ)i=zN/L = ρ(z, θ) and wiji=zN/L,j=z(cid:48)N/L = w(z, z(cid:48)). Then (cid:32) R ≤ ρ(z(cid:48), π)ρ(z(cid:48)(cid:48), θ) sup z∈[z(cid:48)(cid:48)−cL/2,z(cid:48)(cid:48)+cL/2] w(z, z(cid:48)) − w(z(cid:48)(cid:48), z(cid:48)) + z∈[z(cid:48)(cid:48)−cL/2,z(cid:48)(cid:48)+cL/2],z(cid:48)(cid:48)(cid:48)∈[z(cid:48)−cL/2,z(cid:48)+cL/2] sup w(z, z(cid:48)(cid:48)(cid:48)) − w(z, z(cid:48)) (cid:33) (3.15) (3.16) Hence, if we set c = N−α, 0 < α < 1, then as N → ∞, the number of neurons in the local neighborhood cN approaches infinity as N 1−α while c → 0. Then, R → 0 as N → ∞ if limz→z(cid:48)(cid:48) w(z, z(cid:48)) − w(z(cid:48)(cid:48), z(cid:48)) = 0 and limz(cid:48)(cid:48)(cid:48)→z(cid:48) w(z, z(cid:48)(cid:48)(cid:48)) − w(z, z(cid:48)) = 0, i.e. wij is a continuous function in both indices. A similar argument shows that 2(cid:88) k+ cN i=k− cN 2 1 cN (Ii − Ik)ηi(θ) → 0 (3.17) if Ii approaches a continuous function in index i in the infinite N limit. Then (3.4) and (3.5) can be written as ∂θ [1 − cos θ + (Ik + ak) (1 + cos θ)] ρk(θ) = 0 (cid:88) j ak = 2 L N wkjρj(π) (3.18) (3.19) Equations (3.18) and (3.19) form a mean field theory that is realized in a nonuniform coupled network in the infinite size limit as long as the input and coupling function are continuous functions. By linearity, the time dependent mean field theory should equally apply if the external input and the coupling are continuous functions of the indices. continuous, (cid:80) In the N → ∞ limit, setting i → zN/L, ai(t) → a(z, t), ρi(θ, t) → ρ(z, θ, t), Ii → I(z) is Ω dz, and wij → w(z, z(cid:48)) is continuous, we can write mean field i → (N/L)(cid:82) theory in continuum form as ∂tρ(z, θ, t) + ∂θ [1 − cos θ + (I(z) + a(z, t))(1 + cos θ)] ρ(z, θ, t) = 0 w(z, z(cid:48))ρ(z(cid:48), π, t)dz(cid:48) ∂ta(z, t) = −βa(z, t) + 2β (cid:90) The stationary solutions obey ∂θ [1 − cos θ + (I(z) + a(z))(1 + cos θ)] ρ(z, θ) = 0 w(z, z(cid:48))ρ(z(cid:48), π)dz(cid:48) a(z) = 2 (cid:90) 7 (3.20) (3.21) (3.22) FIG. 1. a) Mean field theory synaptic drive for b) connectivity weight w(z) = −J0 +J2 cos(2π/Lz), J0 = 0.2 and J2 = 0.8, and c) external input I(z) = I0 + sin(2π/L(z − z0)), I0 = 1, z0 = 0.25. The stationary solutions will be qualitatively different depending on the sign of I + a. Consider first the suprathreshold regime where I + a > 0. We can then solve (3.21) to obtain (cid:112)I(z) + a(z) w(z, z(cid:48))(cid:112)I(z) + a(z)dz(cid:48) (cid:90) a(z) = 1 π which has been normalized such that(cid:82) ρ(z, θ)dθ = 1. Inserting this back into (3.22) gives π [1 − cos θ + (I(z) + a(z))(1 + cos θ)] ρ(z, θ) = (3.23) (3.24) In the subthreshold regime, I + a < 0, (3.23) has a singularity at 1 − cos θ + (I(z) + a(z))(1 + cos θ) = 0, for which there are two solutions θ± that coalesce in a saddle node bifurcation at I + a = 0. Although ρ is no longer differentiable at equilibrium in the subthreshold regime there is still a weak solution. It has been shown previously [15] that θ− is stable and θ+ is unstable for a single theta neuron. This implies that the density is given by ρ(z, θ) = δ(θ − θ−) and that ρ(z, π) = 0 (i.e. no firing) as expected in the subthreshold regime. Figure 1, shows an example of a stationary "bump" solution for the periodic coupling function, w(z) = −J0 + J2 cos(2π/Lz), which has been used in models of orientation tuning of visual cortex [3] and the rodent head direction system [32]). IV. BEYOND MEAN FIELD THEORY In the infinite N limit when mean field theory applies, the fields η and u are completely described by their means. The time trajectories of these fields are independent of the initial conditions of the individual neurons. For finite N , the trajectories can differ for 8 0.00.20.40.60.81.0z0.20.40.60.81.01.21.4ua)20015010050050100150200i-j6420246810wijb)0.00.20.40.60.81.0z0.40.60.81.01.21.41.6Ic) different initial conditions and going beyond mean field theory involves understanding these fluctuations. Implicit in going beyond mean field theory is that these fields are themselves random variables that are drawn from a distribution functional. In this section, we will derive this distribution functional formally and then use it to compute perturbative expressions for the covariances of η and u. Recall, that the microscopic system is fully described by ∂tηi(θ, t) + ∂θFi(θ, ui)ηi(θ, t) − δ(t − t0)η0 i (θ) = 0 ui(t) + βui(t) − 2β L N wijηj(π, t) − δ(t − t0)u0 i = 0 N(cid:88) j=1 (4.1) (4.2) where we have expressed the initial conditions as forcing terms. The probability density functional for the fields is then comprised of point masses constrained to the dynamical system marginalized over the distribution of the initial data densities: (cid:90) Dη0 (cid:89) i P[η, u] = (cid:34) ×δ ui + βui − 2β L N j δ(cid:2)∂tηi + ∂θFi(θ, ui)ηi − δ(t − t0)η0 i (θ)(cid:3) (cid:35) (cid:88) wijηj(π, t) − δ(t − t0)u0 P[η0] i where P [η0] is the probability density functional of the initial neuron densities for all neurons and Dη0 is the functional integration measure. We consider the initial condition of u to be fixed to u0. Using the functional Fourier transform for the Dirac delta functionals, we then obtain P[η, u] = DηDη0Du e (cid:90) (cid:82) dtdθ ηi[∂tηi+∂θFi(θ,ui)ηi−δ(t−t0)η0 −(cid:80) (cid:82) dt ui[ ui+βui−2β L −(cid:80) × e (cid:80) i i (θ)] j wij ηj (π,t)−δ(t−t0)u0 i ]P[η0] i (4.4) where ηi and ui are response fields for neuron i with functional integration measures Dη and i (θ) = δ(θ − θi(t = 0)), the distribution over initial densities Du over all neurons. If we set η0 is given by the distribution over the initial phase, ρ0 N i (θ). Thus we can write(cid:82) Dη0P [η0] = (cid:80) i dθρ0 i (θ). The initial condition contribution is given by the integral (cid:82)(cid:81) (4.3) (4.5) (4.6) (4.7) eW0[η] = = i (θi)e i ηi(θi,t0) i (θ)e ηi(θ,t0) = e i (θ)(e ηi(θ,t0)−1)) dθiρ0 (cid:90) (cid:89) (cid:90) (cid:89) i ln(1−(cid:82) dθρ0 (cid:80) dθρ0 i i 9 Hence, the system given by (2.6) and (2.7) can be mapped to the distribution functional P[η, u] =(cid:82) DηDη0Due−S with action S = Sη + Su given by (cid:90) π −π (cid:90) t1 (cid:90) t1 t0 (cid:88) (cid:88) i i t0 Sη = Su = (cid:32) ui + βui − 2β (cid:88) j L N dt ui(t) (cid:88) i 1 − (cid:18) (cid:33) (cid:90) (cid:19) i (θ)(e ηi(θ,t0) − 1) dθρ0 (4.8) (4.9) wijηj(π, t) − δ(t − t0)u0 i dt dθ ηi(θ, t)[∂tηi(θ, t) + ∂θF (θ, ui)ηi(θ, t)] + ln The exponential in the initial data contribution to the action (which corresponds to a generating function for a Poisson distribution) can be bilinearized via the Doi-Peliti-Janssen transformation [5 -- 8, 10, 18]: ψi = ηi exp(−ηi), ψi = exp( ηi) − 1, resulting in dθdt ψi(θ, t)[∂tψi(θ, t) + ∂θF (θ, ui)ψi(θ, t)] + (cid:32) ui(t) + βui − δ(t − t0)u0 i − 2β L N dtui(t) (cid:18) (cid:90) 1 − ln dθρ0 i (θ) ψi(θ, t0) (cid:19) (cid:33) (4.10) (4.11) wij( ψj(π, t) + 1)ψj(π, t) (cid:88) (cid:88) i j where we have not included the noncontributing terms that arise after integration by parts. We now make the coarse graining transformation i → zN/L, ui(t) → u(z, t), ψi(θ, t) → Ω dz, and wij → w(z − z(cid:48)), which ψ(z, θ, t), ρi(θ, t) → ρ(z, θ, t), Ii → I(z), (cid:80) i → (N/L)(cid:82) (cid:90) (cid:90) (cid:88) (cid:88) i i Sψ = Su = yields Sψ = Su = (cid:90) (cid:90) N L N L dzdθdt ψ(z, θ, t)[∂tψ(z, θ, t) + ∂θF (θ, u)ψ(z, θ, t)] (cid:90) + N L dz ln (cid:18) (cid:90) (cid:18) 1 − dθρ0(z, θ) ψ(z, θ, t0) (cid:19) (cid:90) (4.12) (cid:19) dzdt u(z, t) u + βu − δ(t − t0)u0(z) − 2β dz(cid:48)w(z − z(cid:48))( ψ(z(cid:48), π, t) + 1)ψ(z(cid:48), π, t) We examine perturbations around the mean field solutions a(z, t) and ρ(z, θ, t) of (3.20) with u → a(z, t)H(t − t0) + v(z, t), u → v, ψ → ρ(z, θ, t)H(t − t0) + ϕ(z, θ, t), and ψ → ϕ, 10 (4.13) where ρ(z, θ, t = t0) = ρ0(z, θ) and H(t − t0) is the Heaviside function. We then obtain dθdtdz ϕ[∂tϕ + ∂θ [1 − cos θ + (I + (a + v))(1 + cos θ)] ϕ + ∂θv(1 + cos θ)ρ dθρ0(z, θ) ϕ(z, θ, t0) dz ln dθdtdz ϕ[∂tϕ + ∂θ [1 − cos θ + (I + (a + v))(1 + cos θ)] ϕ dz + dθ(cid:48) ϕ(z, θ(cid:48), t0)ρ0(z, θ) (cid:19) (cid:90) N L (cid:90) (cid:90) dz dθ(cid:48) ϕ(z, θ(cid:48), t0)ρ0(z, θ) dθ ϕ(z, θ, t0)ρ0(z, θ) (4.14) Sϕ = + (cid:90) (cid:18) 1 − = + ∂θv(1 + cos θ)ρ] − N 2L (cid:90) (cid:90) (cid:90) (cid:90) N L N L N L (cid:90) (cid:20) (cid:90) ( d dt −2β (cid:90) (cid:90) Ω (cid:90) Sϕ = N L (cid:90) Sv = N L Sv = N L dtdz v + β)v − 2β dz(cid:48) w(z − z(cid:48))( ϕ(z(cid:48), π, t) + 1)ϕ(z(cid:48), π, t) dz(cid:48) w(z − z(cid:48)) ϕ(z(cid:48), π, t)ρ(z(cid:48), π) − δ(t − t0)(u0(z) − a(z, t0)) (4.15) Ω We have only included the quadratic term of the initial condition since it is the only one that plays a role at first order perturbation theory (tree level). Finally, if we set the mean (cid:21) field solutions to the stationary solutions ρ(z, θ) and a(z) we obtain dθdtdz ϕ[∂tϕ + ∂θ [1 − cos θ + (I + (a + v))(1 + cos θ)] ϕ + ∂θv(1 + cos θ)ρ] + dz dθ(cid:48) ϕ(z, θ(cid:48), t0)ρ0(z, θ) dθ ϕ(z, θ, t0)ρ0(z, θ) (4.16) (cid:90) (cid:90) N 2L (cid:20) (cid:90) ( d dt −2β (cid:90) Ω (cid:90) (cid:21) dtdz v + β)v − 2β dz(cid:48) w(z − z(cid:48))( ϕ(z(cid:48), π, t) + 1)ϕ(z(cid:48), π, t) dz(cid:48) w(z − z(cid:48)) ϕ(z(cid:48), π, t)ρ(z(cid:48), π) (4.17) Ω Without loss of generality, we set L = 1. In the limit of N → ∞, the dominant term in the probability density functional for the fields will be the extrema of the action, which defines mean field theory. Moments of the fields can be computed perturbatively as an expansion in 1/N by using Laplace's method around mean field (i.e. a loop expansion). The bilinear terms in the action (comprising of a product of a field and a response field) are the linear response functions or propagators. All the other terms are vertices. Each vertex contributes a factor of N while each propagator contributes 1/N . To make the scaling more transparent, we make the rescaling transformation where v → v/N and ϕ → ϕ/N . This change will rescale the propagators to order unity and the vertices to order one or higher depending on how 11 FIG. 2. a) Propagators, ∆ν ν(z, t; z(cid:48), t(cid:48)) (upper left), ∆ϕ ϕ(z, θ, t; z(cid:48), t(cid:48)) ϕ(z, θ, t; z(cid:48), θ(cid:48), t(cid:48)) (lower right) b) Vertices for action in (4.18). From left to ν (z, t; z(cid:48), θ(cid:48), t(cid:48)) (lower left), ∆ν (upper right), and ∆ϕ right, they are ∂θ(1 + cos θ), 0, ρ0(z, θ)ρ0(z, θ), − 2β N w(z − z(cid:48))ρ(z(cid:48), π), − 2β N w(z − z(cid:48)) many response fields they possess. The resulting action is (cid:90) (cid:90) + ∂θv(1 + cos θ)ρ] + dz (cid:20)(cid:18) d (cid:90) 1 2N (cid:19) v − 2β (cid:90) (cid:90) Ω Sv = dtdz v + β dt (cid:90) Ω −2β N dz(cid:48) w(z − z(cid:48)) ϕ(z(cid:48), π, t)ρ(z(cid:48), π) (cid:21) Sϕ = dθdtdz ϕ[∂tϕ + ∂θ [1 − cos θ + (I(z) + (a(z) + v))(1 + cos θ)] ϕ (cid:90) dθ(cid:48) ϕ(z, θ(cid:48), t0)ρ0(z, θ) dθ ϕ(z, θ, t0)ρ0(z, θ) dz(cid:48) w(z − z(cid:48))( ϕ(z(cid:48), π, t)/N + 1)ϕ(z(cid:48), π, t) (4.18) The propagators and vertices can be represented by Feynman graphs or diagrams (see Fig. 2). Each response field corresponds to an outgoing branch (branch on the left) and each field corresponds to an incoming branch (branch on the right). Time flows from right to left and causality is respected by the propagators. To each branch is attached a corresponding propagator. 12 a)b)  ∆v The propagators are defined by v (x; y(cid:48)) ϕ(y; y(cid:48)) v(x; x(cid:48)) ∆ϕ ϕ(y; x(cid:48)) ∆ϕ ∆v G−1 ≡ −1 =  δ2S δ2S (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)v,ϕ,v, ϕ=0 δv(x)δv(x(cid:48)) δ2S δv(x)δϕ(y(cid:48)) δ2S δ ϕ(y)δϕ(y(cid:48)) δ ϕ(y)δv(x(cid:48)) −2βw(z − z(cid:48))δ(π − θ(cid:48))δ(t − t(cid:48)) (d/dt + β)δ(x − x(cid:48))  = ∂θ(1 + cos θ)ρ(z, θ)δ(x − x(cid:48)) (∂t + ∂θ[1 − cos θ + (I + a)(1 + cos θ)])δ(y − y(cid:48)) (4.20) a(x; x(cid:48)) is the response of field a at the nonprimed location to field b at the primed location. The propagator satisfies the condition where x = (z, t), and y = (z, θ, t). The propagator ∆b dq(cid:48)(cid:48)G−1(q, q(cid:48)(cid:48))G(q(cid:48)(cid:48), q(cid:48)) =  δ(x − x(cid:48)) 0  0 δ(y − y(cid:48)) (4.19)  (4.21) (4.22) (4.23) (cid:90) (cid:90) (cid:90) (d/dt + β)∆v where q is x or y as appropriate. Inserting (4.20) into (4.21) yields ϕ(z(cid:48)(cid:48), π, t; x(cid:48)) = δ(x − x(cid:48)) ϕ(z(cid:48)(cid:48), π, t; y(cid:48)) = 0 dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))∆v dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))∆ϕ v(x; x(cid:48)) − 2β v (x; y(cid:48)) − 2β (d/dt + β)∆ϕ (∂t + ∂θ[1 − cos θ + (I(z) + a(z))(1 + cos θ)])∆v ϕ(y; x(cid:48)) + ∂θ(1 + cos θ)ρ(z, θ)∆v v(x; x(cid:48)) = 0 (∂t + ∂θ[1 − cos θ + (I(z) + a(z))(1 + cos θ)])∆ϕ ϕ(y; y(cid:48)) + ∂θ(1 + cos θ)ρ(z, θ)∆ϕ v (x; y(cid:48)) = δ(y − y(cid:48)) (4.24) (4.25) A. Computation of Propagators In order to perform perturbation theory we must compute the Green's functions or prop- agators. There are four types of propagators at each spatial location. The propagator equations are comprised of two sets of 2N coupled integro-partial-differential equations. They can be simplified to ordinary differential equations, which greatly reduces the com- putational complexity. The solutions of the equations change qualitatively depending on whether I + a > 0, suprathreshold regime, and I + a ≤ 0, subthreshold regime. Given that the propagators depend on two coordinates, there are four separate cases. However, the subthreshold neurons are by definition silent so propagators with the second variable in the subthreshold regime are zero, which leaves two cases for the first variable being supra- or sub- threshold. 13 1. Suprathreshold regime In the suprathreshold regime, z ∈ {ζ : I+a(ζ) > 0}, we make the following transformation φ = ϑ>(θ) = 2 tan−1 (cid:112)I(z) + a(z) tan θ 2 (4.26) ϕ> : θ → φ where which obeys dφ dθ = dϑ>(θ) dθ = √ 2 I + a (1 − cos θ) + (I + a)(1 + cos θ) = 2πρ(z, θ) (4.27) where the last equality comes from (3.23). This transformation has the nice property that ϑ>(π) = π. Equations (4.22) and (4.23) transform to (d/dt + β)∆v v(z, t; z(cid:48), t(cid:48)) − 2β (cid:90) > dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))(cid:112)I(z(cid:48)(cid:48)) + a(z(cid:48)(cid:48))∆v (cid:90) dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))(cid:112)I(z(cid:48)(cid:48)) + a(z(cid:48)(cid:48))∆ϕ ϕ(z(cid:48)(cid:48), π, t; z(cid:48), t(cid:48)) = δ(z − z(cid:48))δ(t − t(cid:48)) (4.28) ϕ(z(cid:48)(cid:48), π, t; z(cid:48), θ(cid:48), t(cid:48)) = 0 (d/dt + β)∆ϕ v (z, t; z(cid:48), θ(cid:48), t(cid:48)) − 2β > ϕ(z, φ, t;·) = ∆· where we set ∆· Jacobian of the transformation. ϕ(z, ϑ−1 > (φ), t;·)(dθ/dφ) ≡ ∆· (4.29) ϕ(z, φ, t, ;·), where dθ/dφ is the Equation (4.25) transforms to (∂t + ∂φ(dφ/dθ)[1 − cos θ + (I + a)(1 + cos θ)])∆ϕ ϕ(z, φ, t; z(cid:48), θ(cid:48), t(cid:48))(dφ/dθ) + ∂φ(dφ/dθ)(1 + cos θ)ρ(z, θ)∆ϕ v (z, t; z(cid:48), θ(cid:48), t(cid:48)) = δ(z − z(cid:48))δ(t − t(cid:48))δ(φ − ϑ>(θ(cid:48)))(dφ/dθ) (4.30) Now consider (1 + cos θ)ρ(z, θ) = = = = 1 π 1 π 1 π 1 2π (cid:112)I + a(z)(1 + cos θ) (1 − cos θ) + (I + a(z))(1 + cos θ) √ tan2(θ/2) + (I + a) I + a √ I + a (I + a) tan2(φ/2) + (I + a) (cid:112)I + a(z) 1 + cos φ 14 (4.31) where we have used (4.26) and the tangent half-angle formula tan2 θ 2 = 1 − cos θ 1 + cos θ Inserting (4.31) back into (4.30) gives √ (∂t + 2 I + a)∂φ∆ϕ ϕ(z, φ, t; z(cid:48), θ(cid:48), t(cid:48)) √ − sin φ 2π I + a v (z, t; z(cid:48), θ(cid:48), t(cid:48)) = δ(x − x(cid:48))δ(φ − ϑ>(θ(cid:48)))δ(t − t(cid:48)) ∆ϕ (4.32) (4.33) Similarly, we obtain √ I + a∂φ)∆v (∂t + 2 ϕ(z, φ, t; z(cid:48), t(cid:48)) − sin φ √ v(z, t; z(cid:48), t(cid:48)) = 0 ∆v (4.34) 2π I + a The transformed propagator equations are given by equations (4.28), (4.29), (4.33), and (4.34). Eqs. (4.33) and (4.34) are advection equations in φ, which can be integrated to (cid:90) t (cid:90) t t(cid:48) t(cid:48) ϕ(z, φ, t; z(cid:48), t(cid:48)) = C(z) ∆v ϕ(z, φ, t; z(cid:48), θ(cid:48), t(cid:48)) = C(z) ∆ϕ where dτ sin(φ − ν>(z)(t − τ ))∆v v(z, τ ; z(cid:48), t(cid:48)) dτ sin(φ − ν>(z)(t − τ ))∆ϕ v (z, τ ; z(cid:48), θ(cid:48), t(cid:48)) + δ(φ − ϑ>(θ(cid:48)) − ν>(z)(t − t(cid:48)))δ(z − z(cid:48)) (4.35) C(z) ≡ 2π(cid:112)I(z) + a(z) ν>(z) ≡ 2(cid:112)I(z) + a(z) 1 (cid:90) t We then define the following variables: rv(z, t; z(cid:48), t(cid:48)) = ∆v rϕ(z, t; z(cid:48), θ(cid:48), t(cid:48)) = ∆ϕ dτ sin(ν>(z)(t − τ ))∆v v(z, τ ; z(cid:48), t(cid:48)) ϕ(z, π, t; z(cid:48), t(cid:48)) = C(z) (cid:90) t ϕ(z, π, t; z(cid:48), θ(cid:48), t(cid:48)) − δ(π − ϑ>(θ(cid:48)) − ν>(z)(t − t(cid:48)))δ(z − z(cid:48)) t(cid:48) dτ sin(ν>(z)(t − τ ))∆ϕ v (z, τ ; z(cid:48), θ(cid:48), t(cid:48)) = C(z) t(cid:48) We thus obtain after repeated derivatives and using the propagator equations (4.28), (4.29), 15 (4.36) (4.37) (4.38) (4.33), and (4.34): ( d dt d2 d2 dt2 rv(z, t; z(cid:48), t(cid:48)) = 1 π + β)∆v v(z, t; z(cid:48), t(cid:48)) − β (cid:90) v(z, t; z(cid:48), t(cid:48)) − ν2 ∆v >(z)rv(z, t; z(cid:48), t(cid:48)) (4.39) dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))rv(z(cid:48)(cid:48), t(cid:48)(cid:48); z(cid:48), t(cid:48)) = δ(z − z(cid:48))δ(t − t(cid:48)) (4.40) 1 π dt2 rϕ(z, t; z(cid:48), θ(cid:48), t(cid:48)) = (cid:90) v (z, t; z(cid:48), θ(cid:48), t(cid:48)) − ν2 ∆ϕ v (z, t; z(cid:48), θ(cid:48), t(cid:48)) − β d ( dt = βw(z − z(cid:48))ν>(z(cid:48))δ(π − ϑ>(θ(cid:48)) − ν>(z(cid:48))(t − t(cid:48))) + β)∆ϕ >(z)rϕ(z, t; z(cid:48), θ(cid:48), t(cid:48)) dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))rϕ(z(cid:48)(cid:48), t(cid:48)(cid:48); z(cid:48), θ(cid:48), t(cid:48)) (4.41) (4.42) The covariance function (4.72) involves the integral quantity (cid:90) U (z, t; z(cid:48), t0) = dθ(cid:48)∆ϕ v (z, t; z(cid:48), t0, θ(cid:48))ρ0(z(cid:48), θ(cid:48)) (4.43) by our choice of transformation convention. However, instead of computing the propagator at all values of θ(cid:48), we create another pair of ODEs for U . Applying the integral operator (cid:82) dθ(cid:48)ρ0(z(cid:48), θ(cid:48)) to (4.41) and (4.42) gives d2 1 dt2 r(z, t; z(cid:48), t(cid:48)) = π + β)U (z, t; z(cid:48), , t(cid:48)) − β ( d dt (cid:90) > U (z, t; z(cid:48), t(cid:48)) − ν2 >(z)r(z, t; z(cid:48), t(cid:48)) dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))r(z(cid:48)(cid:48), t(cid:48)(cid:48); z(cid:48), t(cid:48)) (cid:90) dθ(cid:48)ρ0(z(cid:48), θ(cid:48))δ(π − ϑ(θ(cid:48)) − ν>(z(cid:48))(t − t(cid:48))) = βw(z − z(cid:48))ν>(z(cid:48)) = βw(z − z(cid:48))ν>(z(cid:48))ρ0(z(cid:48), θ0) dθ dφ ρ0(z(cid:48), θ0) ρ(z(cid:48), θ0) w(z − z(cid:48))ν>(z(cid:48)) β 2π = (cid:12)(cid:12)(cid:12)(cid:12)θ(cid:48)=θ0 (4.44) (4.45) where r(z, t; z(cid:48), t(cid:48)) =(cid:82) rϕ(z, t; z(cid:48), θ(cid:48), t(cid:48))ρ0(z(cid:48), θ(cid:48)) dθ(cid:48) and θ0 = −ϑ−1(ν>(z(cid:48))(t− t(cid:48)))). Hence, we 16 need to numerically integrate the following equations >(z)r(z, t; z(cid:48), t(cid:48)) U (z, t; z(cid:48), t(cid:48)) − ν2 dt2 r(z, t; z(cid:48), t(cid:48)) = 1 π d2 (cid:90) + β)U (z, t; z(cid:48), , t(cid:48)) − β dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))r(z(cid:48)(cid:48), t(cid:48)(cid:48); z(cid:48), t(cid:48)) = βw(z − z(cid:48))ν>(z(cid:48)) 1 2π (4.47) > (4.46) d2 dt2 rv(z, t; z(cid:48), t(cid:48)) = 1 π + β)∆v v(z, t; z(cid:48), t(cid:48)) − β (cid:90) v(z, t; z(cid:48), t(cid:48)) − ν2 ∆v >(z)rv(z, t; z(cid:48), t(cid:48)) (4.48) dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))rv(z(cid:48)(cid:48), t(cid:48)(cid:48); z(cid:48), t(cid:48)) = δ(z − z(cid:48))δ(t − t(cid:48)) (4.49) dt2 rϕ(z, t; z(cid:48), π, t(cid:48)) = + β)∆ϕ > 1 π (cid:90) v (z, t; z(cid:48), π, t(cid:48)) − ν2 ∆ϕ v (z, t; z(cid:48), π, t(cid:48)) − β ∞(cid:88) > >(z)rϕ(z, t; z(cid:48), π, t(cid:48)) (4.50) dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))rϕ(z(cid:48)(cid:48), t(cid:48)(cid:48); z(cid:48), π, t(cid:48)) β 2 w(z − z(cid:48))δ(t − t(cid:48)) + β w(z − z(cid:48))δ(t − t(cid:48) − Tl(z(cid:48))) l=1 = (4.51) where Tl(z(cid:48)) = {sν(z)s = 2π} marks the time intervals from t(cid:48) such that 2πl−ν>(z(cid:48))Tl(z(cid:48)) = 0. The source at t = t(cid:48) in (4.51) has a factor of one half because because it comes form the θ delta function, which is symmetric about θ = θ(cid:48) since the propagator is symmetric at θ = θ(cid:48), unlike the contribution from the time delta function, which is one sided due to causality. ( d dt ( d dt d2 ( d dt are 2. Subthreshold regime In the subthreshold regime, namely I + a ≤ 0, the mean field solution for the density ρ is a point mass, and this will change the form of the propagators. The propagator equations (d/dt + β)∆v v(x; x(cid:48)) − 2β v (x; y(cid:48)) − 2β dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))∆v dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))∆ϕ ϕ(z(cid:48)(cid:48), π, t; x(cid:48)) = δ(x − x(cid:48)) ϕ(z(cid:48)(cid:48), π, t; y(cid:48)) = 0 (d/dt + β)∆ϕ (∂t + ∂θ[1 − cos θ + (I(z) + a(z))(1 + cos θ)])∆v + ∂θ(1 + cos θ)δ(θ − θ−(z))∆v (∂t + ∂θ[1 − cos θ + (I(z) + a(z))(1 + cos θ)])∆ϕ v (x; y(cid:48)) = δ(y − y(cid:48)) + ∂θ(1 + cos θ)δ(θ − θ−(z))∆ϕ v(x; x(cid:48)) = 0 ϕ(y; x(cid:48)) ϕ(y; y(cid:48)) (4.52) (4.53) (4.54) (4.55) where the equations are defined on I(z) + a(z) < 0, and θ± are the mean field fixed points, where sin θ± = ±2(cid:112)I + a/(1+I +a). However, note that the primed variables are defined (cid:90) (cid:90) 17 over the entire z domain since subthreshold neurons can receive input from suprathreshold neurons. We simplify these equations by breaking the domain of θ into two pieces: D1 = (θ+, θ−) and D2 = (θ−, θ+). In the two advection equations, there will be a clockwise advection of the propagators towards θ− in D1 and in a counterclockwise advection towards θ− in D2. π is in D1 but not D2 so neurons starting in D2 will never fire. In D1, we make the transformation ϑ< : θ → χ: χ = ϑ<(θ) = ln (cid:33) χ = 2 coth−1 (cid:32) tan θ 2 sin θ −(cid:112)I + a(1 + cos θ) sin θ +(cid:112)I + a(1 + cos θ) (cid:112)I(z) + a(z) 2(cid:112)I + a (cid:112)I + a 2(cid:112)I + a (I + a2 + 1) cosh2(χ/2) − 1 (1 − cos θ) − I + a(1 + cos θ) (I + a2 + 1)(cosh(χ) + 1) − 2 1 + cos θ = 2 1 + I + a coth2 χ/2 dχ dθ = dθ dχ = dθ dχ = (cid:90) > (4.56) (4.57) (4.58) (4.59) (4.60) (4.61) (4.62) (4.63) which maps D1 to the real line where −∞ corresponds to θ+ and ∞ corresponds to θ−. We then have the following propagator equations in the χ representation: (d/dt + β)∆v v(z, t; z(cid:48), t(cid:48)) − β dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))∆v ϕ(z(cid:48)(cid:48), π, t; z(cid:48), t(cid:48)) = δ(z − z(cid:48))δ(t − t(cid:48)) (∂t + ν<∂χ)∆v (d/dt + β)∆ϕ (∂t + ν<∂χ)∆ϕ (cid:90) ϕ(z, χ, t; z(cid:48), t(cid:48)) = Q(z, χ)∆v v (z, t; z(cid:48), θ(cid:48), t(cid:48)) − β ϕ(z, χ, t; z(cid:48), θ(cid:48), t(cid:48)) = Q(z, χ)∆ϕ > v(z, t; z(cid:48), t(cid:48)) dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))∆ϕ ϕ(z(cid:48)(cid:48), π, t; z, θ(cid:48), t(cid:48)) = 0 (4.64) v (z, t; z(cid:48), θ(cid:48), t(cid:48)) + δ(z − z(cid:48))δ(χ − ϑ<(θ(cid:48)))δ(t − t(cid:48)) (4.65) 18 where Q(z, χ) = −∂χ and ν<(z) = 2(cid:112)I(z) + a(z). Integrating yields 1 + I + a coth2 χ/2 2 δ(cid:0)ϑ−1(χ) − θ−(z)(cid:1) ϕ(z, χ, t; w(cid:48)) = ∆v ϕ(z, χ, t; w(cid:48)) = ∆ϕ Q(z, χ − ν<(z)(t − τ ))∆v v(z, τ ; z(cid:48), t(cid:48))dτ Q(z, χ − ν<(z)(t − τ ))∆ϕ v (z, τ ; z(cid:48), θ(cid:48), t(cid:48))dτ + δ(z − z(cid:48))δ(χ − ϑs(θ(cid:48)) − ν<(z)(t − t(cid:48))) Hence, the only contribution from the subthreshold neurons are from any neuron that is initially in D1, which for uniformly distributed phases the probability will be (1 − (θ+ − θ−)/2π. The subthreshold propagators are thus passively driven by the superthreshold propagators. Hence, for z in the subthreshold regime, the relevant propagator equations are: (d/dt + β)∆v v(z, t; z(cid:48), t(cid:48)) − β dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))rv(z(cid:48)(cid:48), t; z(cid:48), t(cid:48)) = δ(z − z(cid:48))δ(t − t(cid:48)) (d/dt + β)∆ϕ v (z, t; z(cid:48), π, t(cid:48)) − β = β 2 dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))rϕ(z(cid:48)(cid:48), t; z(cid:48), π, t(cid:48)) w(z − z(cid:48))δ(t − t(cid:48)) + β w(z − z(cid:48))δ(t − t(cid:48) − Tl(z(cid:48))) (4.70) ∞(cid:88) l=1 (cid:90) t (cid:90) t t(cid:48) t(cid:48) (cid:90) > (cid:90) > (cid:90) (4.66) (4.67) (4.68) (4.69) > dz(cid:48)(cid:48)w(z − z(cid:48)(cid:48))ν>(z(cid:48)(cid:48))r(z(cid:48)(cid:48), t; z(cid:48), t(cid:48)) w(z − z(cid:48))ν>(z(cid:48)) (4.71) (d/dt + β)U (z, t; z(cid:48), t(cid:48)) − β = β 2π B. Covariance functions 1. Drive covariance As described previously [10], the covariances between the fields to order 1/N are com- prised of vertices with two outgoing branches. Using the diagrams in Figures 2 and 3, we 19 obtain N(cid:104)δv(z, t)δv(z(cid:48), t(cid:48))(cid:105) = 2β (cid:90) (cid:90) (cid:26)(cid:90) dz1dz2dτ ∆v v(z, t; z1, τ )∆ϕ + (x ↔ x(cid:48)) − dz1 dθ∆ϕ v (z(cid:48), t(cid:48); z2, π, τ )w(z1 − z2)ρ(z2, π) (cid:27) v (z, t; z1, t0, θ)ρ(z1, θ, t0) (cid:90) × dθ(cid:48)∆ϕ v (z(cid:48), t(cid:48); z1, t0, θ(cid:48))ρ(z1, θ(cid:48), t0) (4.72) Evaluating the covariance function in (4.72) requires computing the propagators using the equations derived in the previous section. Our numerical methods for integrating these equations are in the Appendix. We compared the theory to microscopic simulations of (2.4) with fixed initial condition of u(z) set to the mean field solution a(z), and the initial condition of θ(z) is sampled from the probability distribution obeying the mean field solution ρ(z, θ). For the supra-threshold region, the cumulative distribution function for ρ(z, θ) is (cid:32) (cid:33) P (z, θ) = tan−1 1 π (cid:112)I(z) + a(z)(1 + cos θ) sin θ + 1 2 (4.73) from which we can sample θ by applying the inverse of (4.73) to a uniform random number. For the sub-threshold region, all the samples are taken to be at the stable solution θ−(z) = −2 tan−1((cid:112)(I(z) + a(z))). A comparison between the variance of synaptic drive fluctuations for the microscopic simulation as a function of space at a fixed time for two values of N and the theory is shown in figure 4a) for external input and synaptic coupling weight as in figure 1. This is a case where all neurons are in the supra-threshold region. We see that the theory starts to break down for smaller system sizes at the local maxima and minima of the variance. This is expected since the theory is valid to order N−1 in perturbation theory and the maxima and minima are where the effective local population is smallest. Figure 4 b) shows the variance near a maximum as a function of N , showing an accurate prediction after N = 800. The sample size for these microscopic simulations is 5 × 105, and we estimate the error of the variance using bootstrap. The error is of order 10−2. A segment of the spatio- temporal dynamics is shown in figure 5. The theory matches the simulation quite well with the greatest deviation near the maxima and minima. Figure 6 shows the two-time and two-space covariances of the synaptic drive for the same network parameters. The spatial covariance mirrors the coupling function as expected. 20 FIG. 3. Tree level diagrams for a) drive covariance (cid:104)vv(cid:105) and b) rate covariance (cid:104)ϕϕ(cid:105). The lower two diagrams are zero for a) and b). For the upper 3 diagrams in a) and b), the first diagram corresponds to the third term, while the second and third diagrams correspond to the first and second terms of eq. 4.72 and eq. 4.80 respectively. 21 a)b) FIG. 4. Variance times N at time t = 10 for parameters in figure 1. a) Comparison between microscopic simulation and theory calculation for N = 200 and N = 800. b) The N dependence of N(cid:104)δu(z)u(z)(cid:105) at z = 0.2. Standard errors for the microscopic simulation are estimated by bootstrap. Figure 7 shows a comparison between the theory and the microscopic simulation when subthreshold neurons are included. There is a good match when N is large. As N decreases the theory starts to fail at the edges of the bump first. This is likely due to the fact that the location of the edge could move and this is not captured by the theory since it assumes fluctuations around a stationary mean field solution. However, the spontaneous firing of sub-threshold neurons due to either the initial conditions or from the fluctuating inputs of supra-threshold neurons can cause the edge of the bump to move and this is a 22 1.4 1.5 1.6 1.7 1.8 1.9 0 0.5 1N<δuδu>za) 1.78 1.8 1.82 1.84 0 200 400 600 800 1000 1200 1400 1600N<δuδu>Nb) FIG. 5. Spatial-temporal dynamics of the synaptic drive variance for the microscopic simulation for N = 800 in a) and c), and the theory in b) and d). Parameters are as in Figure 1. nonperturbative effect. 2. Rate Covariance The firing rate is defined as ν = 2η(z, π, t) with mean (cid:104)ν(z, π, t)(cid:105) = 2ρ(z, π, t) (4.74) 23 FIG. 6. Spatiotemporal plot of covariance (cid:104)δu(z, 20)δu(z, 20 − τ )(cid:105) for a) theory and b) mi- croscopic simulation using parameters from Figure 1. c) Covariance at a single spatial location, (cid:104)δu(0.5, 20)δu(0.5, 20 − τ )(cid:105). d) Covariance at a single time, (cid:104)δu(z = .005, 20)δu(z(cid:48), 20)(cid:105). Standard errors are estimated by jackknife. and covariance (cid:104)δν(z, t)δν(z(cid:48), t(cid:48))(cid:105) = (cid:104)(ν(z, π, t) − (cid:104)ν(z, π, t)(cid:105))(ν(z(cid:48), π, t(cid:48)) − (cid:104)ν(z(cid:48), π, t(cid:48))(cid:105))(cid:105) = (cid:104)ν(z, π, t)ν(z(cid:48), π, t(cid:48))(cid:105) − (cid:104)ν(z, π, t)(cid:105)(cid:104)ν(z(cid:48), π, t(cid:48))(cid:105) = 4(cid:104)η(z, π, t)η(z(cid:48), π, t(cid:48)) − 4ρ(z, π, t)ρ(z(cid:48), π, t(cid:48))(cid:105) = 4(cid:104)( ϕ(z, t)ϕ(z, t) + ϕ(z, t))( ϕ(z(cid:48), t(cid:48))ϕ(z(cid:48), t(cid:48)) + ϕ(z(cid:48), t(cid:48)))(cid:105) = 4(cid:104)ϕ(z, t)ϕ(z(cid:48), t(cid:48))(cid:105) + 4(cid:104)ϕ(z, t) ϕ(z(cid:48), t(cid:48))(cid:105)(cid:104)ϕ(z(cid:48), t(cid:48))(cid:105) = 4(cid:104)ϕ(z, t)ϕ(z(cid:48), t(cid:48))(cid:105) + ϕ(z, π, t; z(cid:48), π, t(cid:48))ρ(z(cid:48), π, t(cid:48)) ∆ϕ 4 N (4.75) (4.76) (4.77) (4.78) (4.79) At tree level, from the diagrams in Figure 3 b), N(cid:104)ϕ(z, t)ϕ(z(cid:48), t(cid:48))(cid:105) = 2β dz1dz2dτ ∆v ϕ(z, t; z1, τ )∆ϕ (cid:90) (cid:26)(cid:90) + (x ↔ x(cid:48)) − dz1 dθ∆ϕ × dθ(cid:48)∆ϕ ϕ(z(cid:48), t(cid:48); z1, t0, θ(cid:48))ρ(z1, θ(cid:48), t0) ϕ(z(cid:48), π, t(cid:48); z2, π, τ )w(z1 − z2)ρ(z2, π, τ ) (cid:27) ϕ(z, t; z1, t0, θ)ρ(z1, θ, t0) (4.80) (cid:90) (cid:90) 24 -1.6-1.2-0.8-0.4 0 0.4 0.8 1.2 1.6 2 0 0.5 1N<δu(z)δu(z')>z'd) FIG. 7. a) Variance multiplied by N and b) mean of the synaptic drive with subthreshold neurons for constant stimulus I = −1 and c) coupling weight w(z) = A exp(−az)−exp(−bz)+A exp(−a(L− z)) − exp(−b(L − z)) with A = 150, a = 30, and b = 20. The supra-threshold edge of bump is at u = 1. a) is evaluated at time 10. Standard errors are estimated by bootstrap. 25 0 10 20 30 0 0.5 1N<δuδu>zedge of bumpa)-2 0 2 4 6 8 10 0 0.5 1uzb)-10 0 10 20 30 40 50-200-150-100-50 0 50 100 150 200wiji-jc) We rewrite as N(cid:104)ϕ(z, t)ϕ(z(cid:48), t(cid:48))(cid:105) = 2β (cid:90) dz1dz2dτ {rv(z, t; z1, τ )(rϕ(z(cid:48), t(cid:48); z2, π, τ ) (cid:90) (cid:90) (cid:90) (cid:26)(cid:90) (cid:90) (cid:90) (cid:88) (cid:90) l + 2β ν>(z(cid:48)) (cid:90) (cid:90) + (x ↔ x(cid:48)) − − 1 2π − 1 4π2 +δ(π − ϑ>(π) + ν>(z(cid:48))(t(cid:48) − τ ))δ(z(cid:48) − z2))w(z1 − z2)ρ(z2, π, τ )} + (x ↔ x(cid:48)) − dθ(rϕ(z, t; z1, θ, t0) + δ(π − ϑ>(θ) + ν>(z)(t − t0))δ(z − z1))ρ(z1, θ, t0) dz1 × dθ(cid:48)(rϕ(z(cid:48), t(cid:48); z1, θ(cid:48), t0) + δ(π − ϑ>(θ(cid:48)) + ν>(z)(t(cid:48) − t0))δ(z(cid:48) − z1))ρ(z1, θ(cid:48), t0) (cid:27) = 2β dτ dz1dz2rv(z, t; z1, τ )rϕ(z(cid:48), t(cid:48); z2, π, τ )w(z1 − z2)ρ(z2, π, τ ) dz1rv(z, t; z1, t(cid:48) − 2πl/ν)w(z1 − z(cid:48))ρ(z(cid:48), π, τ ) (cid:90) (cid:90) dz1dθrϕ(z, t; z1, θ, t0)ρ(z1, θ, t0) dθ(cid:48)rϕ(z(cid:48), t(cid:48); z1, θ(cid:48), t0)ρ(z1, θ(cid:48), t0) dθrϕ(z, t; z(cid:48), θ, t0)ρ(z(cid:48), θ, t0) − 1 2π dz1δ(z − z1)δ(z(cid:48) − z1) dθ(cid:48)rϕ(z(cid:48), t(cid:48); z, θ(cid:48), t0)ρ(z, θ(cid:48), t0) where r(z, t; z(cid:48), t(cid:48)) = Hence (cid:104)δν(z, t)δν(z(cid:48), t(cid:48))(cid:105) = ν>(z)ν>(z(cid:48)) (cid:90) (cid:18)8β (cid:90) (cid:88) N l ϕ(z, π, t; z(cid:48), t(cid:48)) = rv(z, t; z(cid:48), t(cid:48)) ∆v ϕ(z, π, t; z(cid:48), θ(cid:48), t(cid:48)) = rϕ(z, t; z(cid:48), θ(cid:48), t(cid:48)) + δ(π − ϑ>(θ(cid:48)) + ν>(z)(t − t(cid:48)))δ(z − z(cid:48)) ∆ϕ rϕ(z, t; z(cid:48), θ(cid:48), t(cid:48))ρ0(z(cid:48), θ(cid:48)) dθ(cid:48) (cid:90) (cid:90) dτ dz1dz2rv(z, t; z1, τ )rϕ(z(cid:48), t(cid:48); z2, π, τ )w(z1 − z2)ρ(z2, π, τ ) dz1rv(z, t; z1, t(cid:48) − 2πl/ν>(z(cid:48)))w(z1 − z(cid:48))ρ(z(cid:48), π, τ ) + (x ↔ x(cid:48)) + ν>(z(cid:48))N 8β (cid:90) − 4 N − 2 πN 4 N + dz1r(z, t; z1, t0)r(z(cid:48), t(cid:48); z1, t0) r(z, t; z(cid:48), t0) − 2 πN ϕ(z, π, t; z(cid:48), π, t(cid:48))ρ(z(cid:48), π, t(cid:48)) ∆ϕ r(z(cid:48), t(cid:48); z, t0) − 1 π2N (cid:90) (cid:19) dz1δ(z − z1)δ(z(cid:48) − z1) This quantity is well behaved for t (cid:54)= t(cid:48) and z (cid:54)= z(cid:48). However, in the limit of t(cid:48) → t−, the 26 rate covariance is singular since ϕ(z, π, t; z(cid:48), π, t(cid:48))ρ(z(cid:48), π, t(cid:48)) = δ(z − z(cid:48))δ(ν>(z)(t − t(cid:48))) t(cid:48)→t− ∆ϕ lim (cid:12)(cid:12)(cid:12)(cid:12)θ=π dφ dθ ρ(z(cid:48), π, t(cid:48)) = ν>(z) 2ν>(z) δ(z − z(cid:48))δ(0)ρ(z(cid:48), π, t(cid:48)) (4.81) (4.82) We regularize the singularity at t = t(cid:48) by considering the time integral over a small interval: ∆ν(z, t) = δν(z, s)ds (cid:90) t+∆t/2 (cid:90) t−∆t/2 giving (cid:104)∆ν(z, t)∆ν(z(cid:48), t(cid:48))(cid:105) ∆t2 (cid:90) (cid:18)8β (cid:90) (cid:88) (cid:90) N l = ν>(z)ν>(z(cid:48)) + 8β ν>(z(cid:48))N + (x ↔ x(cid:48)) − 4 N dτ dz1dz2rv(z, t; z1, τ )rϕ(z(cid:48), t(cid:48); z2, π, τ )w(z1 − z2)ρ(z2, π, τ ) dz1rv(z, t; z1, t(cid:48) − 2πl/ν>(z(cid:48)))w(z1 − z(cid:48))ρ(z(cid:48), π, t(cid:48) − 2πl/ν>(z(cid:48))) dz1r(z, t; z1, t0)r(z(cid:48), t(cid:48); z1, t0) (cid:90) (cid:19) dz1δ(z − z1)δ(z(cid:48) − z1) − 2 πN 2 + r(z, t; z(cid:48), t0) − 2 πN ρ(z(cid:48), π, t(cid:48))δ(z − z(cid:48)) N ∆t r(z(cid:48), t(cid:48); z, t0) − 1 π2N (4.83) We regularize the singularity at z = z(cid:48) by taking a local spatial average over [−cN/2 + z, cN/2 + z]. We make the approximation that within this local region, the propagator is constant on space, which is valid under the large N limit. This results in (cid:104)∆ν(z, t)∆ν(z(cid:48), t(cid:48))(cid:105) ∆t2 dτ dz1dz2rv(z, t; z1, τ )rϕ(z(cid:48), t(cid:48); z2, π, τ )w(z1 − z2)ρ(z2, π, τ ) dz1rv(z, t; z1, t(cid:48) − 2πl/ν>(z(cid:48)))w(z1 − z(cid:48))ρ(z(cid:48), π, t(cid:48) − 2πl/ν>(z(cid:48))) (cid:90) (cid:90) (cid:18)8β (cid:90) (cid:88) (cid:90) N l = ν>(z)ν>(z(cid:48)) + 8β ν>(z(cid:48))N + (x ↔ x(cid:48)) − 4 N dz1r(z, t; z1, t0)r(z(cid:48), t(cid:48); z1, t0) (cid:19) − 2 πN r(z, t; z(cid:48), t0) − 2 πN r(z(cid:48), t(cid:48); z, t0) − 1 π2N c + 2 ∆tcN ρ(z(cid:48), π, t(cid:48)) (4.84) Figure 8 shows a comparison of the theory in (4.84) to the microscopic simulations. As shown in Fig. 8 a), at N = 1200, the theory predicts the mean firing rate well. In Fig. 8 c), we show the variance of the firing rate at fixed location. In Fig. 8d), we show the spatial structure of the variance. Again, the theory captures the simulations. 27 FIG. 8. a) Comparison between theory and microscopic simulations of time dependence of mean firing rate at one spatial location. b) Spatial dependence of mean firing rate at time 3. c) and d) show the same comparisons for the variance given in equation (4.84). Parameters are from Fig 1 and N=1200. Standard errors are estimated by bootstrap. V. DISCUSSION Our goal was to understand the dynamics of a large but finite network of determinis- tic synaptically coupled neurons with nonuniform coupling. In particular, we wanted to quantify the dynamics of individual neurons within the network. We first showed that a 28 0.32 0.33 0.34 0.35 0.36 0 1 2 3<υ(z=0.15)>ta) 0.2 0.3 0.4 0.5 0.6 0 0.25 0.5 0.75 1<υ(z)>zb) 0.28 0.29 0.3 0.31 0.32 0 1 2 3<∆υ-(z)∆υ-(z)>/∆t2tc) 0.1 0.2 0.3 0.4 0.5 0.6 0 0.25 0.5 0.75 1<∆υ-(z)∆υ-(z)>/∆t2zd) self-consistent local mean field theory can describe the dynamics of a single network if the external input and coupling weight are continuous functions. This imposes a spatial metric on the network where neurons within a local neighborhood experience similar inputs and can thus be averaged over locally. This local continuity does not impose any conditions on long range interactions, which can still be random. We thus propose a new kind of network to study - continuous randomly coupled spiking networks, where the coupling is continuous but irregular at longer scales. We show that corrections to mean field theory can be computed as an expansion in the number of neurons in a local neighborhood. In this paper, we have chosen to scale the local neighborhood to the total number of neurons but this is not necessary. We do this by first writing down a formal and complete statistical description of the theory, mirroring the Klimontovich approach used in the kinetic theory of plasmas [13, 18, 21]. This formal theory is regularized by averaging, which leads to a BBGKY moment hierarchy. As in previous works [5 -- 8, 10, 18], we showed that the Klimontovich description can be mapped to an equivalent Doi-Peliti-Jansen path integral description from which a perturbation expansion in terms of Feynman diagrams can be derived. The path integral formalism is a convenient tool for calculations. Although we only computed covariances to first order (tree level) it is straight forward (although computationally intensive) to continue to higher order as well as compute higher order moments. We only considered a deterministic network for clarity but our method can easily incorporate stochastic effects, which would just add a new vertex to the action. We showed that the theory works quite well for large enough network size, which can be quite small if all neurons receive suprathreshold input. However, the expansion works less well for neurons with critical input such as neurons at the edge of a bump where infinitesimally small perturbations can produce qualitatively different behavior. Quanti- tatively capturing the dynamics at the edge may require renormalization. The formalism could be a systematic means to understanding randomly connected networks [28] and the so-called balanced network [27, 31], where the mean inputs from excitatory and inhibitory synapses are attracted to a fixed point near zero and the neuron dynamics is dominated by the fluctuations. 29 VI. ACKNOWLEDGMENTS This research was supported by the Intramural Research Program of the NIH, NIDDK. VII. APPENDIX: NUMERICAL METHODS 1. Discretization schemes We use full backward's Euler for green function calculation for propagators. = sij drij dt Uij − ν2 dsij dt ∂tUij = −βUij + 1 π = i rij β N (cid:88) j wijνjrjk + β 2π wijνj ij + hst ij 1 π ij = rt−1 rt ij = st−1 st ij = U t−1 U t ij + h( ij + h(−βU t ij − ν2 i rt U t ij) β N ij + (cid:88) j wijνjrt jk + β 2π wijνj) ij ij − hst ij = rt−1 rt (cid:88) ij − h 1 i rt st ij + hν2 π ij − h β U t N j ij = st−1 U t ij wijνjrt jk + hβU t ij = U t−1 ij + h β 2π wijνj  I hν2. ∗ I −hβ/N w. ∗ ν(cid:48) 0 −hI I −h/πI I + hβI 0   rt·j st·j U t·j  =  rt−1·j st−1·j U t−1 ·j + hβ/2πw.jνj  (7.1) (7.2) (7.3) (7.4) (7.5) (7.6) (7.7) (7.8) (7.9) (7.10) We add the spike terms in equation (4.51) directly to the propagator ∆ϕ ν (z, t; z(cid:48), π, t(cid:48)) for all possible l when t− Tl(z(cid:48)) = t(cid:48). These spike terms add stiffness to the differential equation 30 and explicit differential equation solvers like Runge-Kutta have poor stability properties. [1] L. F. Abbott and Carl van Vreeswijk. Asynchronous states in networks of pulse-coupled oscillators. Phys. Rev. E, 48:1483 -- 1490, Aug 1993. [2] Shun-ichi Amari. Dynamics of pattern formation in lateral-inhibition type neural fields. Bio- logical Cybernetics, 27(2):77 -- 87, Jun 1977. [3] Sompolinsky H. Ben-Yishai R, Bar-Or RL. Theory of orientation tuning in visual cortex. Proc Natl Acad Sci U S A., 1995. [4] Nicolas Brunel and Vincent Hakim. Fast global oscillations in networks of integrate-and-fire neurons with low firing rates. Neural Computation, 11(7):1621 -- 1671, 1999. [5] Michael Buice and Carson Chow. Generalized activity equations for spiking neural network dynamics. Frontiers in Computational Neuroscience, 7:162, 2013. [6] Michael A. Buice and Carson C. Chow. Correlations, fluctuations, and stability of a finite-size network of coupled oscillators. Phys. Rev. E, 76:031118, Sep 2007. [7] Michael A Buice and Carson C Chow. Beyond mean field theory: statistical field theory for neural networks. Journal of Statistical Mechanics: Theory and Experiment, 2013(03):P03003, 2013. [8] Michael A. Buice, Jack D. Cowan, and Carson C. Chow. Systematic fluctuation expansion for neural network activity equations. Neural Computation, 22(2):377 -- 426, 2010. PMID: 19852585. [9] Rashmi C. Desai and Robert Zwanzig. Statistical mechanics of a nonlinear stochastic model, volume 19. 07 1978. [10] Michael A.Buice Carson C.Chow. Dynamics finite size effects in spiking neural network. PLOS, January 2013. [11] Michael A. Cohen and Stephen Grossberg. Absolute stability of global pattern formation and parallel memory storage by competitive neural networks. In Stephen Grossberg, editor, The Adaptive Brain I, volume 42 of Advances in Psychology, pages 288 -- 308. North-Holland, 1987. [12] David Dahmen, Hannah Bos, and Moritz Helias. Correlated fluctuations in strongly coupled binary networks beyond equilibrium. Phys. Rev. X, 6:031024, Aug 2016. [13] D.R.Nicholson. Introduction to Plasma Theory, volume 2. Cambridge University Press, 1984. 31 [14] Grgory Dumont, Alexandre Payeur, and Andr Longtin. A stochastic-field description of finite- size spiking neural networks. PLOS Computational Biology, 13(8):1 -- 34, 08 2017. [15] B. Ermentrout. Type i membranes, phase resetting curves, and synchrony. Neural-Comput, 8, 1996. [16] Nicolas Fourcaud and Nicolas Brunel. Dynamics of the firing probability of noisy integrate- and-fire neurons. Neural Computation, 14(9):2057 -- 2110, 2002. [17] M Helias, T Tetzlaff, and M Diesmann. Echoes in correlated neural systems. New Journal of Physics, 15(2):023002, 2013. [18] Eric J. Hildebrand, Michael A. Buice, and Carson C. Chow. Kinetic theory of coupled oscil- lators. Phys. Rev. Lett., 98:054101, Jan 2007. [19] J J Hopfield. Neural networks and physical systems with emergent collective computational abilities. Proceedings of the National Academy of Sciences, 79(8):2554 -- 2558, 1982. [20] Wilson HR and Cowan JD. Excitatory and inhibitory interactons in localized populations of model neurons. Biophysical Journal, 1:1 -- 24, Jan 1972. [21] S. Ichimaru. Basic Principles of Plasma Physics: A Statistical Approach. A lecture note and reprint series. W. A. Benjamin, 1973. [22] S.R. Jones and N. Kopell. Local network parameters can affect inter-network phase lags in central pattern generators. Journal of Mathematical Biology, 52(1):115 -- 140, Jan 2006. [23] Eva Lang and Wilhelm Stannat. Finite-size effects on traveling wave solutions to neural field equations. Journal of Mathematical Neuroscience, 7:5, 2017. [24] Selva K. Maran and Carmen C. Canavier. Using phase resetting to predict 1:1 and 2:2 locking in two neuron networks in which firing order is not always preserved. Journal of Computational Neuroscience, 24(1):37 -- 55, Feb 2008. [25] Renato E. Mirollo and Steven H. Strogatz. Synchronization of pulse-coupled biological oscil- lators. SIAM J. Appl. Math., 50(6):1645 -- 1662, November 1990. [26] Duane Q. Nykamp and Daniel Tranchina. A population density approach that facilitates large- scale modeling of neural networks: Extension to slow inhibitory synapses. Neural Computation, 13(3):511 -- 546, 2001. [27] Srdjan Ostojic. Two types of asynchronous activity in networks of excitatory and inhibitory spiking neurons. Nature Neuroscience, 17:594 EP -- , 02 2014. [28] H. Sompolinsky, A. Crisanti, and H. J. Sommers. Chaos in random neural networks. Phys. 32 Rev. Lett., 61:259 -- 262, Jul 1988. [29] Jonathan D. Touboul and G. Bard Ermentrout. Finite-size and correlation-induced effects in mean-field dynamics. Journal of Computational Neuroscience, 31(3):453 -- 484, Nov 2011. [30] Alessandro Treves. Mean-field analysis of neuronal spike dynamics. Network: Computation in Neural Systems, 4(3):259 -- 284, 1993. [31] C van Vreeswijk and H Sompolinsky. Chaos in neuronal networks with balanced excitatory and inhibitory activity. Science, 274:1724 -- 1726, 1996. [32] K Zhang. Representation of spatial orientation by the intrinsic dynamics of the head-direction cell ensemble: a theory. Journal of Neuroscience, 16(6):2112 -- 2126, 1996. 33
1202.2491
1
1202
2012-02-12T06:22:12
Analysis of inverse stochastic resonance and the long-term firing of Hodgkin-Huxley neurons with Gaussian white noise
[ "q-bio.NC" ]
In previous articles we have investigated the firing properties of the standard Hodgkin-Huxley (HH) systems of ordinary and partial differential equations in response to input currents composed of a drift (mean) and additive Gaussian white noise. For certain values of the mean current, as the noise amplitude increased from zero, the firing rate exhibited a minimum and this phenomenon was called inverse stochastic resonance (ISR). Here we analyse the underlying transitions from a stable equilibrium point to the limit cycle and vice-versa. Focusing on the case of a mean input current density $\mu=6.8$ at which repetitive firing occurs and ISR had been found to be pronounced, some of the properties of the corresponding stable equilibrium point are found. A linearized approximation around this point has oscillatory solutions from whose maxima spikes tend to occur. A one dimensional diffusion is also constructed for small noise based on the correlations between the pairs of HH variables and the small magnitudes of the fluctuations in two of them. Properties of the basin of attraction of the limit cycle (spike) are investigated heuristically and also the nature of distribution of spikes at very small noise corresponding to trajectories which do not ever enter the basin of attraction of the equilibrium point. Long term trials of duration 500000 ms are carried out for values of the noise parameter $\sigma$ from 0 to 2.0, with results appearing in Section 3. The graph of mean spike count versus $\sigma$ is divided into 4 regions $R_1,...,R_4,$ where $R_3$ contains the minimum associated with ISR.
q-bio.NC
q-bio
Analysis of inverse stochastic resonance and the long-term firing of Hodgkin-Huxley neurons with Gaussian white noise Henry C. Tuckwell1†, Jurgen Jost 2 1 Max Planck Institute for Mathematics in the Sciences Inselstr. 22, 04103 Leipzig, Germany † Corresponding author: [email protected] August 17, 2018 2 1 0 2 b e F 2 1 ] . C N o i b - q [ 1 v 1 9 4 2 . 2 0 2 1 : v i X r a 1 Abstract In previous articles we have investigated the firing properties of the standard Hodgkin-Huxley (HH) systems of ordinary and partial differen- tial equations in response to input currents composed of a drift (mean) and additive Gaussian white noise. For certain values of the mean cur- rent, as the noise amplitude increased from zero, the firing rate exhibited a minimum and this phenomenon was called inverse stochastic resonance (ISR). Here we analyse the underlying transitions from a stable equilib- rium point to the limit cycle and vice-versa. Focusing on the case of a mean input current density µ = 6.8 at which repetitive firing occurs and ISR had been found to be pronounced, some of the properties of the corre- sponding stable equilibrium point are found. A linearized approximation around this point has oscillatory solutions from whose maxima spikes tend to occur. A one dimensional diffusion is also constructed for small noise based on the correlations between the pairs of HH variables and the small magnitudes of the fluctuations in two of them. Properties of the basin of attraction of the limit cycle (spike) are investigated heuristically and also the nature of distribution of spikes at very small noise correspond- ing to trajectories which do not ever enter the basin of attraction of the equilibrium point. Long term trials of duration 500000 ms are carried out for values of the noise parameter σ from 0 to 2.0, with results appearing in Section 3. The graph of mean spike count versus σ is divided into 4 regions R1, ..., R4, where R3 contains the minimum associated with ISR. In R1 noise has practically no effect until a critical value of σ = σc1 is reached. At a larger critical value σ = σc2 , the probability of transitions from the basin of attraction of the equilibrium point to that of the limit cycle becomes greater than zero and the spike rate thereafter increases with increasing σ. The quantitative scheme underlying the ISR curve is outlined in terms of exit time random variables and illustrated dia- grammatically. In the final subsection 3.4, several statistical properties of the main random variables associated with long term spiking activity are given, including distributions of exit times from the two relevant basins of attraction and the interspike interval. Short Title: Hodgkin-Huxley Keywords and Phrases: Hodgkin-Huxley equations, noise 1 Introduction The Hodgkin-Huxley [1] systems of ordinary and partial differential equations, based on the electrophysiology of the squid giant axon, are the cornerstone of mathematical models of single neurons as well as several types of cardiac cells. Recent such studies include those of Komendantov et al. [2] for hypothathalamic magnocellular neuroendocrine cells, Saarinen et al. [3] for cerebellar granule cells, Williams et al. [4] for ventricular myocytes, Kameneva et al. [5] for retinal ganglion cells and Drion et al. [6] for dopaminergic neurons. Many of these computational cell models contain 10 or more components as the important roles of many different ion channels have been discovered since the appearance 2 of the HH model. Analysis of such higher-dimensional models is very complex as there may be 50 or more parameters in distinction to the relatively few in the 4-component Hodgkin-Huxley system. Since the latter does in fact embrace some of the basic firing properties of neurons in general, there has naturally been a large number of analyses and computational studies of the HH systems. These include both deterministic (for example, [7, 8, 9, 10, 11, 12]) and stochastic (for example, [13, 14, 15, 16, 17]) modeling. In recent articles [19, 20, 21] we have explored the effects of both additive Gaussian white noise and conductance noise on repetitive firing in the Hodgkin- Huxley system. In the additive noise case for the ordinary differential equation (ODE) model, when the mean input current density µ is not far above the threshold of 6.4 µA/cm2 for repetitive firing, the number of spikes in the first 500 or 1000 ms was found to undergo a pronounced minimum (ISR) as the noise level σ increased from zero [19]. The minimum occurred around σ = 0.35. Guo [22] has recently found similar results for the HH ODE system with colored (Ornstein-Uhlenbeck process) noise. Similar results were found for the partial differential equation (PDE) system [20, 21] where the spatial distribution of the noise was also an important factor, which led to the disinction between the effects of noise on the instigation and propagation of spikes. 1.1 Model description In this article we restrict attention to the HH ODE system, which corresponds to a uniformly polarized or "space-clamped" neuron. The system of stochastic differential equations was given in our previous articles but are repeated here for completeness and notation: dV = 1 C [µ + gKn4(VK − V ) + gN am3h(VN a − V ) + gL(VL − V )]dt + σdW (1) and for the auxiliary variables dn = [αn(1 − n) − βnn]dt dm = [αm(1 − m) − βmm]dt dh = [αh(1 − h) − βhh]dt (2) (3) (4) where C is the membrane capacitance per unit area, µ, which may depend on t, is the mean input current density, gK, gN a and gL are the maximal (constant) potassium, sodium and leak conductances per unit area with corresponding equilibrium potentials VK, VN a, and Vl, respectively. The noise enters as the derivative of a standard Wiener process W and has amplitude σ. The auxiliary variables are n(t), the potassium activation, m(t), the sodium activation and h(t), the sodium inactivation. The coefficients in the differential equations for the auxiliary variables as functions of depolarization are αn(V ) = 10 − V 100[e(10−V )/10 − 1] (5) 3 βn(V ) = e−V /80 1 8 25 − V αm(V ) = 10[e(25−V )/10 − 1] βm(V ) = 4e−V /18 e−V /20 αh(V ) = 7 100 βh(V ) = 1 e(30−V )/10 + 1 (6) (7) (8) (9) (10) 2 Stable equilibrium point and limit cycle When µ is above the critical value µc1 for repetitive firing (saddle-node bifur- cation) and smaller than the value µc2 at which there is a subcritical Hopf bifurcation, there are two attractors consisting of a limit-cycle (action potential trajectory) and a stable equilibrium point. 2.1 Stable equilibrium point x∗ Let us denote the random vector (V, n, m, h) by X = (X1, X2, X3, X4) and rewrite the system of equations as dX1 = F1(X)dt + σdW dXk = Fk(X)dt (11) (12) where k = 2, 3, 4. Because a mean input current density of µ =6.8 µA/cm2 was found to exhibit a pronounced minimum in firing as σ increased, we will focus on this value of µ unless stated otherwise. Setting, with zero noise, F (x∗) = 0 (13) yields equilibrium points for the deterministic Hodgkin-Huxley system. The Jacobian matrix at the equilibrium point is defined as ∂xj x∗ 4 (cid:26) ∂Fi (cid:27) J(x∗) = , i, j = 1, ..., 4. (14) Figure 1: Eigenvalues in the complex λ-plane for three values of µ. The values of the real eigenvalue near λ = −0.13 are indistinguishable in this diagram. Figure 1 shows the eigenvalues in the complex lambda plane for values of µ = 5, below the critical value µc1, µ = 7.5, between the critical values µc1, and µc2 , and µ = 10 which is above µc2. The nature of the critical point at the various values of µ can be readily seen. For the chosen value of µ = 6.8, numerical evaluation gives an equilibrium point at x∗ = (4.0536, 0.38107, 0.084327, 0.45129) (15) at which the components of F are (−0.00000003084, 0.000024246,−0.0000013509, 0.0000048941) (16) Numerical evaluation of the eigenvalues of J(x∗) gives λ1 = −4.641, the complex conjugate pair λ2,3 = −0.630 ± 0.548i and λ4 = −0.1323. Hence x∗ is an asymptotically stable spiral point. 2.2 The linear approximation The system of stochastic ordinary differential equations for the process X ob- tained by linearizing about x∗ is 4(cid:88) k=1 d X1 = Xkdt + σdW (17) ∂F1 ∂xk 5 4(cid:88) d Xj = k=1 ∂Fj ∂xk Xkdt, j = 2, 3, 4, (18) where the partial derivatives are evaluated at x∗. The Jacobian is found nu- merically to be  −0.010891 −1.2764 0.000030551 −0.0019202 0.00032794 −0.000044773 0 0 102 ∗  (19) 1.2710 0.079467 0 −0.034876 0 0 0 −0.0012664 Note that the system (17)-(18) is linear and so X is a Gaussian process. Using the theory in Rodriguez and Tuckwell [23], in the absence of an imposed spiking threshold, the exact distribution of X can be found at any t. Between spikes the linearized system (17)-(18) is expected to provide a rea- sonable approximation to the fully nonlinear system (1)-(4). This is clearly demonstrated by the two sets of sample paths shown in Figure 2. With input parameters µ = 6.8, σ = 0.6, a time segment of length about 60 ms gave the sample path for V (X1) shown in the top part of the Figure. With the same path for the Wiener process (or the white noise), the sample path for the voltage X1 in the linearized system, with the same initial values, is seen in the lower part of Figure 2, to mimic closely that in the upper part. However, there is one very striking difference betwen the two paths as at about t = 59 ms the voltage in the linear system attains a local maximum and then decreases quite rapidly, continuing in an oscillatory fashion. In contrast, at about the same t and volt- age values, a spike arises in the nonlinear system. However, these paths indicate that the time of spiking in the nonlinear system can be well approximated as the time at which the voltage in the linear sytem first attains a threshold value. Such a first passage time can be determined from the usual first exit time theory for diffusion processes. 6 Figure 2: In the top part, voltage is plotted versus time for a sample path in the nonlinear full Hodgkin.Huxley system with the parameters shown. A spike forms near the end of the record. In the lower record is shown the voltage path for the system linearized about the stable equilibrium point. The Wiener path is the same in both records to enable a comparison to be made. 2.3 Numerical examples Without an imposed threshold condition, no spikes are possible in the linearized stochastic system described by (17) and (18) because there is only a stable equi- librium point about which trajectories fluctuate. This is of course in distinction to the Hodgkin-Huxley system (with suitable input parameters) where trajec- tories may, if the fluctuations are large enough, give rise to spikes around the limit cycle. It is of interest to examine some statistical properties of the original process X in nonspiking periods. An example of paths with µ = 6.8 and σ = 0.1 over a 50 ms nonspiking period is shown in Figure 2. The basic statistical prop- erties of the components over a 500 ms time period are given in Table 1. In Table 2 are given the corresponding correlation coefficients. 7 Table 1: Statistics of variables during non-spiking period. µ = 6.8, σ = 0.1 Mean Max 4.446 0.3828 0.0874 0.4542 V 4.049 n 0.3811 m 0.0843 h 0.4515 Min 3.658 0.3795 0.0811 0.4488 St Dev Coef Var 0.1337 6.61e-4 0.0012 0.001 0.0330 0.0017 0.0143 0.0023 Table 2: Correlation coefficients during non-spiking period. µ = 6.8, σ = 0.1 V V 1.0000 n 0.3068 m 0.9562 h -0.2137 n 0.3068 1.0000 0.4649 -0.9894 m 0.9562 0.4649 1.0000 -0.3699 h -0.2137 -0.9894 -0.3699 1.0000 Figure 3: Sample paths for the 4 components of X over a 50 ms period during which there were no spikes. Input parameters µ = 6.8 and σ = 0.1. Tables 3 and 4 give the statistical properties during a 500 ms nonspiking time period when the noise level is σ = 0.6. 8 Table 3: Statistics of variables during non-spiking period. µ = 6.8, σ = 0.6 Mean Max 8.962 0.419 0.1393 0.4668 V 4.149 n 0.383 m 0.0859 h 0.4471 Min -0.206 0.368 0.0538 0.3883 St Dev Coef Var 1.301 0.0077 0.0126 0.0122 0.314 0.020 0.147 0.0274 Table 4: Correlation coefficients, µ = 6.8, σ = 0.6 V V 1.0000 n 0.3505 m 0.9690 h -0.2408 n 0.3505 1.0000 0.5284 -0.9854 m 0.9690 0.5284 1.0000 -0.4272 h -0.2408 -0.9854 -0.4272 1.0000 There are two very noticeable features of the sample paths in the nonspiking periods examined. Firstly, the oscillatory character of the paths, which is trace- able to the eigenvalues of the Jacobian at the equilibrium point. Secondly, the strong positive correlation between V and m and the strong negative correlation between n and h. Figure 4: Showing how single, double and multiple spikes arise from oscillations in the nonlinear full Hodgkin-Huxley system with noise. µ = 6.8 9 The oscillatory nature of the paths can never be captured in the standard integrate and fire models nor the leaky integrate and fire models [24]. When the noise is sufficiently large to make for fairly frequent spiking, the times of spiking must tend to arise at the maxima in the oscillations of V . This is seen dramatically in Figure 3 and in Figure 2 of the previous subsection. 2.4 A one-dimensional diffusion Examination of the statistical properties of the process as given in Tables 1 and 2 shows that for small noise the coefficients of variation of n and h are an order of magnitude smaller than those for V and m and that the correlation coefficient of V and m is close to unity. These observations make it reasonable to consider a 1-dimensional approximation V to the HH system in which a constant n ≈ n and a constant h ≈ h, with m = kV , where k is another constant. These approximations lead, on putting C = 1 in (1), to the following stochastic differential equation for V , d V = (µ + c1 − c2 V + c3 V 3 − c4 V 4)dt + σdW, where c1 = gKn4VK + gLVL c2 = gKn4 + gL c3 = gN ak3hVN a c4 = gN ak3h. (20) (21) (22) (23) (24) Standard theory gives for such a diffusion that the mean exit time from a value x ∈ (a, b) to outside this interval satisfies the ordinary differential equation (σ2/2)M(cid:48)(cid:48) + (µ + c1 − cx + c3x3 − c4 x)M(cid:48) = −1, x ∈ (a, b), (25) where primes denote differentiation, along with suitable boundary conditions. Preliminary investigations of the validity of this approximation were made using values for µ = 6.8 and σ = 0.1 (see Tables 1 and 2) but a detailed study will be reported in a subsequent article. 2.5 The limit cycle and its basin of attraction With constant input current density in the interval [µc1 , µc2), repetitive spiking may occur with a fixed period. For µ = 6.8 and no noise the period is about 17.65 ms. The limit cycle is in 4-space but we here show in the upper part of Figure 2 the projection of the limit cycle obtained by plotting n versus V . In the lower part of the figure is shown the position of the stable equilibrium point, here designated R, for the same value of µ. It can be seen that the limit cycle approaches quite close to R. In a previous article in which the moment method was used to explore the effects of noise on HH spiking (Tuckwell and Jost, 2009), we heuristically estimated the basin of attraction of the stable equilibrium point. 10 Part of the basin of attraction of the limit cycle can be numerically estimated by taking the union of all stochastic paths which do not collapse to the stable equilibrium point. As explained later, there are ranges of values of σ where the stochastic paths have this property and Figure 6 shows a sample of such paths for various such σ. Here paths were taken over 1000 msec involving about 50 spikes for values of σ ranging from 0.05 to 0.25. This picture gives an idea of how far off the deterministic limit cycle a path may wander without entering the basin of attraction of the stable equilibrium point. The variability of paths is relatively small. Figure 5: In the top part, voltage is plotted versus potassium activation vari- ablefor the deterministic path of repetitive spikes, depicting the limit cycle for the HH ODE system with µ = 6.8 > µc1. In the lower part the limit cycle is magnified in the vicinity of the stable rest point. 11 Figure 6: The union of many stochastic paths for values of the noise parameter σ for which paths did not collapse into the stable point in a time interval of at least 1000 ms. Another way to see the limited variability of the times to complete a spike orbit (limit cycle) is displayed in Figure 7. Here spikes were observed during repetitive firing during periods in which no transitions to the basin of attraction of the rest point occurred, with noise levels from σ = 0 to σ = 0.4. The statistical properties of the interspike interval (ISI) are shown in the Figure. The most salient feature is that the mean ISI is practically constant (blue triangles) as the noise varies, staying in the interval [17.58, 17.82] for these values of σ. Naturally the standard deviation of the ISI increases (green squares) as σ increases, the maximum ISI increases /red circles) and the minimum ISI decreases (black diamonds). 12 Figure 7: Some statistical measures of the ISI during repetitive spiking at various noise levels with µ = 6.8. For σ = 0, there is no variability. The mean ISI is shown with (blue) triangles, the (green) squares denote the mean + 3 standard deviations, the (red) circles denote the maxima and the (black) diamonds denote the minima. Until σ is about 0.07, there are no cessations of spiking up to 500000 ms. Examples of distributions of the ISI during repetitive spiking at small noise levels are shown in Figure 8. In the top panel is an ISI histogram for a noise level (σ = 0.07) at which there is apparently no cessation of spiking over extremely long (infinite?) time periods with 28429 ISIs in 500000 ms. The distribution is roughly Gaussian and has a mean of 17.59 ms and a standard deviation of 0.221 ms. For the second histogram shown, σ = 0.085, which is just greater than the critical value at which spiking may stop in a finite time. The number of spikes is only 4397 and the mean and standard deviation of the ISI are 17.60 ms and 0.276 ms, respectively. The distribution of the ISI is practically the same in both cases. 13 Figure 8: Histograms of the ISI during repetitive spiking at two small noise levels. Top, σ = 0.070; bottom, σ = 0.085. For σ = 0.07 there are no cessations of spiking up to 500000 ms whereas for σ = 0.085 firing stops after about 4400 spikes. 3 Results on spiking for the HH system and in- verse stochastic resonance In the following it is assumed that µ is such that repetitive spiking does in fact occur. In relation to the stochastic paths for the HH system of stochastic equations we define the following two random variables. Firstly, the exit time of the process to escape from the basin of attraction BL of the limit cycle L to that BR of the rest point R is by TL→R(x0), x0 ∈ BL (26) where (x0) = (V0, n0, m0, h0) is an initial point. Secondly the exit time for the process to escape from the basin of attraction BR of R to that of the limit cycle is TR→L(x0), x0 ∈ BR. (27) 14 Using the standard theory for diffusion processes, Kolmogorov second order par- tial differential equations for the moments and distributions of these quantities as a function of initial values were described in our previous work (Tuckwell et al., 2009). However, any attempt to solve these equations analytically or even by numerical methods for PDEs, apart from being a formidable task, requires an exact or even approximate knowledge of the two basins of attraction which is unfortunately not presently available. In fact the probabilistic nature of BL and BR is not completely understood because it seems that for small enough noise, at least for some values of µ, escape from BL may be impossible, not being observed in extremely long time periods and similarly, for particular ranges of σ, for escape from BR. 3.1 Long term trials and data collected We are interested in obtaining samples of meaningful sizes for the above random variables and determining their properties, mainly as a function of σ. Since some of the exits from one basin of attraction to the other are, for some values of σ, very rare events the simulations of solutions of the full stochastic HH system have to be performed over very long time periods. These were chosen to be 500000 ms or about 8 minutes and 20 seconds with a timestep of 0.065 ms. Results for the 500000 ms interval were obtained as the union of 100 trials of length 5000 ms, where the final values of V, n, m, h from any trial were used as the initial values for the next trial. Then the data on t, V, n, m, h for the 100 trials were concatenated to give 5 single vectors each with over 7.5 million elements. The records for V were analyzed to determine the times of occurrence of spikes and from these the interspike intervals were determined. In total, 50 trials of length 500000 were simulated. 3.2 All spikes The numbers of spikes were recorded in each of 50 trials of length 500000 ms for 40 values of σ. The mean number of spikes, denoted by E[NSP], is plotted against σ in Figure 9. In the top panel all results are included, whereas in the lower left panel detail is shown for 0 ≤ σ ≤ 0.1 and in the lower right panel, for 0.1 ≤ σ ≤ 0.5. The general form of the plot of E[NSP] versus σ is similar to that in Figure 5 of Tuckwell et al. (2009) where the total time period was much less at 1000 ms. Thus, Figure 9 exhibits the phenomenon of inverse stochastic resonance, which refers to a firing rate which, as noise level increases, at first declines to a minimum and then becomes greater. However, Figure 9 shows more detail and reveals 4 distinct regimes marked R1,...,R4. In region R1, which extends from the deterministic setting of σ = 0 to very close to σ = 0.07, the noise is almost without effect and the number of spikes is always close to that for the zero noise, 28431. Region R2 is characterized by a rapidly falling number of spikes as σ increases from 0.07 to 0.14. At the latter value of σ the mean number of spikes is 104.8 which is about 0.4% of the maximum value. In region 15 R3, from σ = 0.15 to σ = 0.35, E[NSP] stays below 100 with a broad minimum of about 9.5 spikes (0.03% of the maximum) around σ = 0.295 to 0.300. By σ = 0.375, the beginning of region R4, the mean number of spikes has reached over 120 and then increases sharply and eventually at a slower rate to reach 25883 around σ = 2. Figure 9: The dependence on the noise parameter σ of the mean of the total number of spikes E[NSP] in a time interval of length 500000 ms in the nonlinear HH system. Here µ = 6.8 with 50 trials at each point. The values of σ are divided into 4 regimes designated R1 to R4. In the top panel, results are given for all values of σ. In the lower left panel the detail of R1 and the start of R2 are shown. In the lower right panel, the detail of R3 and the start of R4 are shown. 3.3 Underlying scheme for ISR The following observations constitute a basis for ISR which occurs at some values of µ in the HH system described by Equations (1)-(4). We define two critical values σc1 and σc2 of the noise parameter σ, with 0 < σc1 < σc2 < ∞. σc1 and σc2 depend on µ, and are only relevant above the critical value for repetitive firing and possibly for selected initial values of the process (see Tuckwell et al., 2009). With reference to the two random variables defined by (26) and (27), 16 but without any specific values of x0, we have the following, as schematized diagrammatically in Figure 10. 0 < σ < σc1. Repetitive spiking continues, presumably indefinitely. The probability of a transition from BL to BR at any time is zero. The expectation of TL→R(x0), x0 ∈ BL, is infinite. σc1 < σ < σc2. Repetitive spiking ceases at some time t1 < ∞. The probability of a transition from BL to BR is greater than zero and increases with noise level. The expectation of TL→R(x0), x0 ∈ BL, is finite and and tends to decrease with increasing σ. After a transition from BL to BR, the probability of the reverse transition from BR to BL is zero. The expectation of TR→L(x0), x0 ∈ BR, is infinite. After the cessation of repetitive spiking, there is no more spiking. σ > σc2. After a transition from BL to BR, the probability of a transition from BR to BL is greater than zero and tends to increase as σ incresases. The expectation of TR→L(x0), x0 ∈ BR, is finite and tends to decrease as σ increases. After a transition back from BR to BL, the reverse transition occurs with non- zero pobability, followed again by a transition from BL to BR and so forth, indefinitely. Epochs of spiking and nonspiking alternate with variable durations which tend to become shorter as σ increases. Figure 10 illustrates some of these aspects. The bottom panel shows part of the actual E[NSP] curve (normalized) from Figure 9 and the approximate values of σc1 and σc2 are obtained from this curve. In the top panel, the probabilities of transitions from BL to BR and BR to BL are indicated as commencing to increase from zero at the two critical values of σ and rising to unity, although this final value depends on the initial value x0 which is not fixed throughout. In the middle panel, descriptions of the expectations of the exit times from BL and BR are given. 17 Figure 10: A schematic representation for the transitions underlying ISR. There are shown two critical values of the noise parameter σ and the probabilities of transitions between the basins of attraction BL and BR of the limit cycle and rest point are sketched in the top panel for the various ranges of values of σ. In the bottom panel is shown the normalized expected number of spikes obtained by simulation over 500000 ms for µ = 6.8, a value of µ at which ISR had been found to be pronounced. The scheme given in Figure 10 is in accordance with the results for sample paths of V for simulated spike trains shown in Figure 11. Here the time periods are all 500000 ms but in the top two records time is only shown to 2000 ms. The very small noise case gives apparently indefinite repetitive spiking. Somewhat larger noise leads to an initial isolated burst followed by silence, possibly for an infinite time. Increasing the noise further leads to occasional bursting and eventually the frequent bursting case occurs. 18 Figure 11: Illustrating the 4 basic patterns of spiking activity. Top left: σ = 0.05. Incessant spiking, at least to 500000 ms. Top right: σ = 0.25. One initial burst with no further spikes, at least to 500000 ms. Bottom left: σ = 0.34. Occasional bursts with separations of order 100000 ms. Bottom right: σ = 0.45. Frequent bursts. 3.4 Some statistics of TL→R and TR→L In this subsection results on certain statistical aspects of the underlying random variables TL→R and TR→L, defined by (26) and (27), obtained from the long- term simulation of the HH system (1)-(4) are given. 3.4.1 Escape time from BL During repetitive spiking the trajectories of the process do not deviate much from the deterministic limit cycle as illustrated in Figure 6. It is assumed that there is a well-defined but unkown set in (V, n, m, h)-space, denoted by BL, containing the deterministic repetitive spiking trajectory, which is called the basin of attraction of the limit cycle. This implies that if the process started in BL, then, with no noise, the path would approach the limit cycle. With noise, trajectories may escape from BL whereupon they enter the basin of attraction 19 BR of the stable equilibrium point. The nature of BL is unknown and it is not clear that it is a regular open set because if it was it is likely that escape from it would occur eventually with probability one no matter how small the noise level. Thus, even though it seems that for σ < σc1 , P (L → R) = 0 and E[TL→R = ∞], it may be the case that P (L → R) is so small for small enough σ that the event of this escape is unlikely to be observed in the course of feasible simulations. It is a remaining mathematical challenge to ascertain the veracity of these remarks. Notwithstanding these uncertainties, the mean value of E[TL→R] of the exit time from BL was estimated from sample paths and the results are shown for σ ≥ 0.2 in Figure 12. For these noise levels the mean exit time declines rapidly from very large values to a minimum of about 57 ms at σ = 1.25. Thereafter, E[TL→R] seems to increase slightly to about 72 ms at σ = 2. Figure 12: The dependence on the noise parameter σ of the expectation E[TL→R] of the random variable which is the time of exit of the process from the basin of attraction of the limit cycle to that of the stable rest point. 50 trials at each point with µ = 6.8. For values of σ less than a critical value σc1 ≈ 0.07 the value of this expectation is apparently infinite. Extremely large values which occurred for σc1 < σ < 0.2 are not shown. 20 3.4.2 Escape time from BR The basin of attraction of the stable rest point is also unknown exactly. Figures 13 and 14 show estimates of the mean and standard deviation of TR→L over various ranges of values of σ. Again, it is not known with certainty, but it appears that the probability of escape from BR to BL is zero (or extremely close to zero) until σ ≥ σc2. In Figure 13, the mean is shown dropping from values of order 300000 ms at σ = 0.25 to eventually reach values about 30 ms at σ = 2. Figure 14 shows corresponding results for the standard deviation, but only for σ ≥ 0.35 because sample sizes were too small for lesser values of the noise parameter. In the lower panel of Figure 14 is shown the dependence of the coefficient of variation of TR→L on noise level. It can be seen that there is an initial increase in this quantity until a maximum is attained at about σ = 0.5, whereupon it declines monotonically. Figure 13: The dependence on the noise parameter σ of the expectation E[TR→L] of the random variable which is the time of exit of the process from the basin of attraction of the stable rest point to that of the limit cycle. Here µ = 6.8. 50 trials at each point. For values of σ less than some value just less than 0.25 the value of this expectation is apparently infinite. 21 Figure 14: Top two panels. The dependence on the noise parameter σ of the standard deviation σTR→L,of the random variable which is the time of exit of the process from the basin of attraction of the stable rest point to that of the limit cycle. Here µ = 6.8. 50 trials at each point. Bottom panel. The coefficient of variation of the random variable TR→L as a function of σ. 3.4.3 Distributions There are three main random variables of interest in relation to the empirical spike trains obtained in this study. These are the previously defined exit times TL→R and TR→L and in addition the general ISI for the whole train. Samples for the latter random variable can be viewed as the union of those of the previous two. Examples of histograms of these random variables are shown in Figure 15. In all cases the results for 50 trials of duration 500000 ms are combined. In Figure 15A is given a histogram for all ISIs for σ = 0.2. This occurs in region R3 of Figure 9, where there are very few spikes, all with short ISIs as there are no returns to BL fron BR. The histogram is of the same nature as those in Figure 8. See also Figure 7. In Figure 15C the raw histogram is shown for all ISIs with σ = 0.350. There is a preponderance of small intervals and then an exponential-type distribution of longer (> 21.5 ms) intervals as depicted in Figure 15E. The tail of the expo- 22 nential distribution is very long and extends out to about 500000 ms which is the limit in these simulations. The distribution depicted in Figure 15E is close to the actual distribution of TR→L for this value of σ. Similar remarks apply to the pairs of Figures 15B and 15D (σ = 0.375) and 15F and 15G (σ = 0.5). In the final plot of Figure 15H, the number of spikes has become enormous and the majority of ISIs are less than 50 ms as expected for σ = 1 which is in region R4 of Figure 9. In the histograms of Figure 15 C, B, F and H, there are small bin counts out to very large times and these are not visible compared to the large bin count at short intervals. However, they are visible at relatively small values in the truncated histograms. Figure 15: Examples of distributions (histograms) of ISIs and exit times from BR based on 50 trials (data pooled) of length 500000 ms. Red histograms, raw data. Blue histograms, ISIs truncated at 21.5 ms. A. Raw data for σ = 0.2. B. Raw data for σ = 0.375. C. Raw data for σ = 0.350. D. Truncated data for σ = 0.375. E. Truncated data for σ = 0.350. F. Raw data for σ = 0.5. G. Truncated data for σ = 0.5. H. Raw data for σ = 1.0. To complete the picture for the distributions of the three key random vari- ables, Figure 16 shows histograms of mean numbers of spikes per burst for two 23 values, 0.4 and 1.0, of σ. If these are multiplied by the mean ISI within bursts, which is about 17.6 ms, the mean times spent in the basin of attraction of the limit cycle (BL) and hence of TL→R are estimated. For smaller σ some very large means were obtained (not shown). Noticeable in Figure 16 is the smaller magnitude and small variability when σ = 1.0 compared to σ = 0.4. The distributions in both cases shown are approximately Gaussian which may be compared with the exponential-types shown in Figures 15 D, E and G. Figure 16: Histograms of mean numbers of spikes per burst, which indicate magnitudes of exit times from BL to BR. based on 50 trials of length 500000. 4 Discussion The inhibitory effects of noise on repetitive spiking in squid axon, on which the HH system is based, were well documented experimentally by Paydarfar et al. [25]. The first theoretical evidence of ISR in the full HH sytem was obtained for µ = 5 in the non-repetitive spiking mode [26] (Figure 2), as it was found that there was a maximum in the mean ISI at small values of σ. Subsequently ISR has been demonstrated for repetitive spiking in both the HH system of ODEs and PDEs [18, 20, 21, 22]. In addition to using the standard initial conditions and additive noise, in [18] the phenomenon was confirmed with respect to random 24 initial conditions and conductance-based input. Some theory of ISI was also outlined [18] in terms of bifurcations in the HH sytem [7, 10, 27] and of the exit times [28] from the basins of attraction of a stable equilibrium point and a limit cycle. However the simulations were over relatively short time periods and the statistical details of the exit times were not explored in detail. In this article we have first explored the properties of the stable equilibrium point in depth. Linearizing the stochastic HH system about that point gave an approximate sytem of stochastic differential equations whose oscillatory solu- tions were found to mimic those of the full system in nonspiking periods. The oscillations are evidently an integral part of firing in the HH system with the considered parameters, because spikes did tend to emerge at the maxima. The distribution of ISIs with trajectories not departing greatly from the limit cycle for very small noise was estimated and is, as perhaps expected, Gaussian-like. Long-term simulations, to 500000 ms were performed with µ = 6.8 for many values of the noise parameter σ. Four ranges of values of σ were distinguished, based on mean total numbers of spikes. These regions were denoted R1, ..., R4 in Figure 9. Two critical values of σ were also apparent, denoted by 0 < σc1 < σc2 < ∞. A detailed discussion of the mechanisms underlying ISR in terms of these critical values was given in Section 3.3. It is not certain whether escape from the basin of attraction of the limit cycle can ever occur when σ < σc1 or whether escape from the basin of attraction from the basin of attraction of the stable equilibrium point can ever occur when σc1 < σ < σc2, but these events were never observed in 50 trials of length 500000 ms with several values of σ. References [1] A.L. Hodgkin, A.F. Huxley, A quantitative description of membrane cur- rent and its application to conduction and excitation in nerve, J. Physiol. 117 (1952) 500-544. [2] A.O. Komendantov, N.A.Trayanova, J.G. Tasker, Somato-dendritic mechanisms underlying the electrophysiological properties of hypotha- lamic magnocellular neuroendocrine cells: A multicompartmental model study. J. Comput. Neurosci. 23 (2007) 143-168. [3] A. Saarinen, M-L.Linne, O. Yli-Harja, Stochastic differential equation model for cerebellar granule cell excitability. PLoS Comp. Biol. 4 (2008) e1000004. [4] G.S.B. Williams, G.D. Smith, E.A. Sobie, M.S. Jafri MS, Models of cardiac excitation-contraction coupling in ventricular myocytes. Math. Biosci. 226 (2010) 1-15. [5] T. Kameneva, H. Meffin, A.N. Burkitt, Modelling intrinsic electrophysi- ological properties of ON and OFF retinal ganglion cells. J. Comp. Neu- rosci. 31 (2011) 547-561. 25 [6] G. Drion, L. Massotte, R. Sepulchre, V. Seutin V, How modeling can rec- oncile apparently discrepant experimental results: the case of pacemaking in dopaminergic neurons. PLoS Comput. Biol. 7 (2011) e1002050. [7] B. Hassard, Bifurcation of periodic solutions of the Hodgkin- Huxley model for the squid giant axon. J. Theor. Biol. 71 (1978) 401-420. [8] E.N. Best, Null space in the Hodgkin-Huxley equations a critical test. Biophys J 27 (1979) 87-104. [9] K. Aihara, G. Matsumoto G, Two stable steady states in the Hodgkin- Huxley axons. Biophys J 41 (1983) 87-89. [10] H. Fukai, S. Doi, T. Nomura, S. Sato, Hopf bifurcations in multiple- parameter space of the Hodgkin-Huxley equations I. Global organization of bistable periodic solutions. Biol. Cybern. 82 (2000) 215-222. [11] J.Guckenheimer, R.A. Oliva, Chaos in the HodgkinHuxley model. SIAM J. App. Dyn. Sys. 1 (2002)105-114. [12] D. Calitoiu, B.J. Oommen, D. Nussbaum, Spikes annihilation in the Hodgkin-Huxley neuron. Biol. Cybern. 98 (2008) 239-257. [13] Y. Horikawa, Noise effects on spike propagation in the stochastic Hodgkin- Huxley models. Biol. Cybern. 66 (1991)19-25. [14] D. Brown, J-F. Feng, S. Feerick, Variability of firing of Hodgkin- Hux- ley and FitzHugh-Nagumo neurons with stochastic synaptic input. Phys. Rev. Lett. 82 (1999) 4731-4734. [15] P.H.E. Tiesinga, J.V. Jos´e, T.J. Sejnowski, Comparison of current-driven and conductance-driven neocortical model neurons with HodgkinHuxley voltage-gated channels. Phys. Rev. E 62 (2000) 8413-8419. [16] T.D. Austin, The emergence of the deterministic Hodgkin-Huxley equa- tions as a limit from the underlying stochastic ion-channel mechanism. Ann. Appl. Prob. 18 (2006) 1279-1325. [17] M. Ozer, L.J. Graham LJ, Impact of network activity on noise delayed spiking for a Hodgkin-Huxley model. Eur. Phys. J. B 61 (2008) 499-503. [18] H.C. Tuckwell, J.Jost, B.S. Gutkin, Inhibition and modulation of rhyth- mic neuronal spiking by noise. Phys. Rev. E 80 (2009) 031907. [19] H.C. Tuckwell, J. Jost J (2009) Moment analysis of the Hodgkin-Huxley system with additive noise. Physica A 388 (2009) 4115-4125. [20] H.C.Tuckwell, J. Jost, Weak noise powerfully inhibits rhythmic spiking but not its propagation. PLoS Comp. Biol. 6 (2010) e1000794. 26 [21] H.C. Tuckwell, J. Jost, The effects of various spatial distributions of weak noise on rhythmic spiking. J. Comp. Neurosci. 30 (2011) 361-371. [22] D. Guo, Inhibition of rhythmic spiking by colored noise in neural systems. Cogn Neurodyn. 5 (2011) 293-300. [23] R. Rodriguez, H.C. Tuckwell, Statistical properties of stochastic nonlinear dynamical models of single spiking neurons and neural networks. Phys. Rev. E 54(1996) 5585-5590. [24] H.C. Tuckwell. Introduction to Theoretical Neurobiology, Volume 2. Cambridge University Press, Cambridge UK, 1988. [25] D. Paydarfar, D.B. Forger, J.R. Clay , Noisy inputs and the induction of on-off switching behavior in a neuronal pacemaker. J. Neurophysiol. 96 (2006) 3338-3348. [26] H.C. Tuckwell, Spike trains in a stochastic Hodgkin-Huxley system. BioSystems 80 (2005) 25-36. [27] J. Jost, Dynamical Systems, Springer, Berlin, 2005. [28] H. C. Tuckwell, Stochastic Processes in the Neurosciences, SIAM, Philadelphia, 1989. 27
1807.04691
1
1807
2018-07-12T15:57:20
Spatial Embedding Imposes Constraints on the Network Architectures of Neural Systems
[ "q-bio.NC" ]
A fundamental understanding of the network architecture of the brain is necessary for the further development of theories explicating circuit function. Perhaps as a derivative of its initial application to abstract informational systems, network science provides many methods and summary statistics that address the network's topological characteristics with little or no thought to its physical instantiation. Recent progress has capitalized on quantitative tools from network science to parsimoniously describe and predict neural activity and connectivity across multiple spatial and temporal scales. Yet, for embedded systems, physical laws can directly constrain processes of network growth, development, and function, and an appreciation of those physical laws is therefore critical to an understanding of the system. Recent evidence demonstrates that the constraints imposed by the physical shape of the brain, and by the mechanical forces at play in its development, have marked effects on the observed network topology and function. Here, we review the rules imposed by space on the development of neural networks and show that these rules give rise to a specific set of complex topologies. We present evidence that these fundamental wiring rules affect the repertoire of neural dynamics that can emerge from the system, and thereby inform our understanding of network dysfunction in disease. We also discuss several computational tools, mathematical models, and algorithms that have proven useful in delineating the effects of spatial embedding on a given networked system and are important considerations for addressing future problems in network neuroscience. Finally, we outline several open questions regarding the network architectures that support circuit function, the answers to which will require a thorough and honest appraisal of the role of physical space in brain network anatomy and physiology.
q-bio.NC
q-bio
Spatial Embedding Imposes Constraints on the Network Architectures of Neural Systems Jennifer Stiso1,2 and Danielle S. Bassett1,3,4,5,6 1Department of Bioengineering, School of Engineering and Applied Sciences, University of Pennsylvania, Philadelphia, PA, 19104 5Department of Physics & Astronomy, College of Arts and Sciences, University of Pennsylvania, Philadelphia, PA, 19104 3Department of Electrical and Systems Engineering, School of Engineering and Applied Sciences, University of Pennsylvania, Philadelphia, PA, 19104 4Department of Neurology, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA, 19104 2Neuroscience Graduate Group, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA, 19104 and 6To whom correspondence should be addressed: [email protected] A fundamental understanding of the network architecture of the brain is necessary for the further development of theories explicating circuit function. Recent progress has capitalized on quantitative tools from network science to parsimoniously describe and predict neural activity and connectiv- ity across multiple spatial and temporal scales. Perhaps as a historical artifact of its origins in mathematics or perhaps as a derivative of its initial application to abstract informational systems, network science provides many methods and summary statistics that address the networks topologi- cal characteristics with little or no thought to its physical instantiation. Yet, for embedded systems, physical laws can directly constrain processes of network growth, development, and function, and an appreciation of those physical laws is therefore critical to an understanding of the system. Re- cent evidence demonstrates that the constraints imposed by the physical shape and volume of the brain, and by the mechanical forces at play in its development, have marked effects on the observed network topology and function. Here, we review the rules imposed by space on the development of neural networks and show that these rules give rise to a specific set of complex topologies. We present evidence that these fundamental wiring rules affect the repertoire of neural dynamics that can emerge from the system, and thereby inform our understanding of network dysfunction in dis- ease. We also discuss several computational tools, mathematical models, and algorithms that have proven useful in delineating the effects of spatial embedding on a given networked system and are important considerations for addressing future problems in network neuroscience. Finally, we out- line several open questions regarding the network architectures that support circuit function, the answers to which will require a thorough and honest appraisal of the role of physical space in brain network anatomy and physiology. NETWORK TOPOLOGY VERSUS GEOMETRY by studies of regional activation (Box 1). IN NEURAL SYSTEMS In contemporary neuroscience, increasing volumes of data are being brought to bear on the question of how heterogeneous and distributed interactions between neu- ral units might give rise to complex behaviors. Such interactions form characteristic patterns across multiple spatial scales, spanning from the relatively small scales of molecules and cells, to the relatively large scales of areas and lobes [1]. A natural language in which to describe such interactions is network science, which elegantly rep- resents interconnected systems as sets of nodes linked by edges [2]. Intuitively, nodes often represent proteins, neu- rons, subcortical nuclei, or large cortical areas, and edges often represent either (i) structural links in the form of chemical bonds, synapses, or white-matter tracts, or (ii) functional links in the form of statistical relations be- tween nodal activity time series. Generally, the resultant network architecture can be fruitfully studied using tools from graph theory to obtain mechanistic insights perti- nent to cognition [3, 4], above and beyond those provided In particular, several fundamental questions in neuro- science are quintessentially network questions concern- ing the physical relationships between functional units. How does the physical structure of a circuit affect its function? How does coordinated activity at small spa- tial scales give rise to emergent phenomena at large spa- tial scales? How might alterations in neurodevelopmental processes lead to circuit malfunction in psychiatric disor- ders? How might pathology spread through cortical and subcortical tissue giving rise to the well-known clinical presentations of neurological disease? These questions collectively highlight the fact that the brain - and its multiple networks of interacting units - is physically em- bedded into a fixed three-dimensional enclosure. Natural consequences of this embedding include diverse physical drivers of early connection formation and physical con- straints on the resultant adult network architecture. An understanding of the system's constitution and basal dy- namics therefore require not only approaches to quantify and predict network topology, but also tools, theories, and methods to quantify and predict network geometry and its role in both enabling and constraining system function. In this review, we provide evidence to support the no- tion that a consideration of the brain's physical embed- ding will prove critical for a holistic understanding of neu- ral circuit function. We focus our comments on the utility of informing this consideration with emerging computa- tional tools developed for the characterization of spatial networks. Indeed, while network science was originally developed in the context of systems devoid of clear spatial characteristics [5], the field has steadily developed tools and intuitions for spatially embedded network systems [6]. In the light of these recent technical developments, we begin by recounting observations from empirical stud- ies addressing the question of how brain networks are embedded into physical space. Next, we discuss the rele- vance of this spatial embedding for an understanding of network function and dysfunction. We complement these empirical discussions with a more technical exposition on the relevant tools, methods, and statistical approaches to be considered when analyzing brain networks. Lastly, we close by outlining a few important future directions in methodological development and neuroscientific investi- gation that would benefit from an honest appraisal of the role of space in neural network architecture and dynam- ics. PHYSICAL CONSTRAINTS ON NETWORK TOPOLOGY AND GEOMETRY The processes and influences that guide the formation of structural connections in neural systems are disparate and varied [7, 8]. Evidence from genetics suggests that neurons with similar functions as operationalized by sim- ilar gene expression tend to have more similar connection profiles than neurons with less similar functions [7, 9, 10]. Of course, it is important to note that some spatial sim- ilarity of expression profiles is expected due to the in- fluence of small scale spatial gradients in growth factors over periods of development [7]. However, evidence from the Allen Brain Atlas suggests that interareal connec- tivity profiles in rodent brains are even more correlated with gene co-expression than expected simply based on such spatial relationships [9]. This heightened correlation could be partially explained by observations in mathe- matical modeling studies that neurons with similar in- puts (and therefore potentially performing similar func- tions) tend to have more similar connection profiles than neurons with dissimilar inputs [11]. Yet, while genetic coding and functional utility each play important roles, a key challenge lies in summariz- ing the various constraints on connection formation in a simple and intuitive theory that can guide future pre- dictions. One particularly acclaimed candidate mecha- nism for such a theory is that of physical constraints on the development, maintenance, and use of connections. Metabolism related to neural architecture and function 2 is costly, utilizing 20% of the body's energy, despite com- prising only 2% of its volume [12, 13]. Even the devel- opment of axons alone, comprising only a small portion of cortical tissue, extorts a large material cost [7]. The existence of these pervasive costs motivated early work to postulate that wiring minimization is a fundamental driver of connection formation. Consistent with this hy- pothesis, the connectomes of multiple species are pre- dominantly comprised of wires extending over markedly short distances [13–19], and this observation holds across different methods of data collection [16, 18, 19]. However, mounting evidence suggests that pressures for wiring minimization may compete against pressures for efficient communication [20–22]. Early evidence sup- porting the role of efficient communication came from the observation that one can fix the network architecture of inter-areal projections in the macaque cortex and then rearrange the location of areas in space to obtain a con- figuration with significantly (32%) lower wiring cost than that present in the real system [21]. A similar method can be used to obtain a configuration of the C. elegans neuronal connectome with 48% lower wiring cost than that present in the real system [21]. Interestingly, the connections whose length is decreased most also tend to be those that shorten the characteristic path length – one of many ways to quantify how efficiently a network can communicate [21, 23]. Consistent observations have been made in human white matter tractography [20], the mouse inter-areal connectome [10], and dendritic arbors [7, 14]. Notably, computational models that instantiate both constraints on wiring and efficient communication produce topologies more similar to the true topologies than models that instantiate a constraint on wiring min- imization alone [11, 24]. Moreover, models that allow for changes in this tradeoff over developmental time peri- ods better fit observed connectome growth patterns than prior models [25]. It is precisely this balance between wiring minimiza- tion and communication efficiency that is thought to pro- duce the complex network topologies observed in neural systems, along with markedly precise spatial embedding [24, 26]. To better understand this spatially embedded topology, it is useful to consider methods that can si- multaneously (rather than independently) assess topol- ogy and geometry. One such method that has proven particularly useful in the study of neural systems from mice to humans is Rentian scaling, which assesses the efficiency of a network's spatial embedding [20, 27–30]. Originally developed in the context of computer circuits, Rentian scaling describes a power-law scaling relation- ship between the number of nodes in a volume and the number of connections crossing the boundary of the vol- ume [7, 20]. The existence of such a power law relation- ship with an exponent known as Rent's exponent is con- sistent with an efficient spatial embedding of a complex topology [31, 32]. In turn that efficient spatial embedding is thought to support a broad repertoire of functional dynamics. For example, tracts that bridge disparate ar- eas of cortex to increase communication efficiency despite greater wiring cost, also critically add to the functional diversity of the brain in a manner that is distinct from that predicted by path length alone [33]. REFLECTIONS OF PHYSICAL CONSTRAINTS IN LOCAL, MESOSCALE, AND GLOBAL NETWORK TOPOLOGY Across species, the brain consistently exhibits a set of topological features at local, meso-, and global scales that can be relatively simply explained by spatial wiring rules [34–36]. At the local scale, multiple modalities have been used to demonstrate that a key conserved topologi- cal feature is the existence of hubs, or nodes of unexpect- edly high degree [37, 38]. Such hubs emerge naturally in computational models in which the location of nodes are fixed in space, and edges between nodes are rewired to miminize average wiring length and to maximize topo- logical efficiency by minimizing the average shortest path length (Box 1). Interestingly, the number and degree of hubs varies systematically with the relative importance of the two constraints [14, 24] (Fig. 1). When wiring min- imization is not enforced, networks become star graphs with a single giant hub [14, 24]; when both constraints are balanced, networks contain several hubs of varying degrees, consistent with the topology observed in brain networks [24]. It is notable that such constraints can be implemented within the natural processes of devel- opment; for example, in adult C. elegans, hub neurons have been tracked back to the earliest born neurons in the embryo, which accumulate a large number of connec- tions along the normative growth trajectory [26, 39]. At the mesoscale, a key conserved topological feature is modularity, or the existence of internally dense and exter- nally sparse communities of nodes [34, 40]. The strength of modularity in a network is commonly quantified using a modularity quality index (Box 1). In computational models, this index obtained under pressures of wiring minimization and communication efficiency (quantified with path length) was more similar to that empirically measured in the connectomes of the macaque and C. el- egans than to that obtained under either constraint sep- arately [8, 24]. Again it is notable that such constraints can be implemented within the natural processes of de- velopment; for example, in C. elegans, communities form when many neurons are born in a similar temporal win- dow, and therefore typically share a common progenitor type, spatial location, and genetic profile [26, 41]. Spaces between modules can form cavities or cycles, or intu- itively holes in the network, that can be identified with emerging tools from applied algebraic topology (Box 2) [42]. The locations, prevalence, and weight structure of these cycles differs markedly between geometric and ran- dom networks [43, 44], with patterns of functional con- nectivity among neurons exhibiting characteristics simi- lar to those observed in spatially constrained geometric 3 networks [45]. It will be interesting in future to gain a deeper understanding of the relations between cycles and modules, and their emergence through the spatially constrained processes of development. At the global scale, a key conserved topological feature is small-worldness, or the confluence of unexpectedly high clustering and short path length (Box 1) [46]. Such an architecture is thought to be particularly conducive to a balance between local information processing within the clusters, and global information transmission across the topologically long distance connections [47]. Similar to the existence of hubs, modules, and cavities, small-world architecture in a network can naturally arise from spa- tial constraints on wiring [48]. Intuitively, clusters tend to form in spatially nearby regions in order to minimize wiring cost, while long distance connections facilitating efficient communication tend to form only occasionally due to their elevated wiring cost [49]. In concert with these empirical observations, computational models that account for wiring economy produce networks with small- world architecture reminiscent of that observed in real neural systems [50]. Collectively, these studies demon- strate the influence of parsimonious wiring rules on com- plex network topology. Future work could be directed to better understand the aspects of connectome topology that remain unexplained and thus may arise from more subtle rules [8]. RELEVANCE OF NETWORK GEOMETRY FOR DYNAMICS AND COGNITION Pressures for wiring minimization and communica- tion efficiency can exist alongside developmental pro- cesses that produce non-isotropically structured organs. Such processes include migration, elongation, segrega- tion, folding, and closure that accompany neurulation resulting in the bilaterally symmetric nervous system composed of proencephalon, mesencephalon, rhomben- cephalon, and spinal cord as well as patterning across multiple overlapping signaling gradients. It is intuitively possible that such processes could also explain the ob- served differences in the network topologies of different sectors of the brain [51, 52], which can impinge on the functions that those sectors are optimized to perform (Box 3). Indeed, prior work has noted the co-existence of complex structural topologies and spatial gradients of specific function [53], although it has been difficult to achieve a mechanistic understanding of exactly how the two relate to one another. One particularly promising recent line of investigation has proposed the existence of a set of primary spatial gradients that explain variance in large-scale connectivity [54, 55]. In both humans and macaques, the primary axis of variance is bounded on one end by the transmodal default mode system, and on the other end by the unimodal sensory systems [55]. No- tably, this gradient is tightly linked to the geometry of the network, with the regions located at one end having 4 FIG. 1. Effect of wiring minimization and communication efficiency on network topology. Networks were generated by modulating the balance between a constraint on wiring (in the figure referred to as a spatial cost) and a constraint on information routing efficiency (in the figure referred to as a temporal cost). The parameter β, which ranges between 0 and 1, tunes this balance by weighting spatial cost against temporal cost. When β = 0 only the spatial cost is considered, while when β = 1 only the temporal cost is considered. (A) Examples of networks at different values of β when only the spatial constraint exists (left), when only the temporal constraint exists (right), and when the two constraints are balanced (middle). Root nodes are shown in green and all other nodes are shown in yellow. (B) Spatial costs (blue) and temporal costs (red) vary as a function of β. This figure was adapted with permission from [14]. maximal spatial distances from the regions located at the other end [55]. Additionally, the regions located at the peaks of the transmodal gradient have substantial overlap with structural hubs in human connectomes [4, 56]. As discussed earlier, these hub could serve to optimally bal- ance wiring cost and communication efficiency across the connectome, in addition to explaining patterns of func- tional connectivity. Put simply, such evidence supports the notion that the cortex is fundamentally organized along a dimension of function from concrete to abstract, and that dimension manifests clearly in the network's spatial embedding. The diverse spatial location of cortical and subcortical areas has important implications for the patterns of neu- ral dynamics that one would expect to observe. Consider, for example, the patterns of intrinsic activity noted con- sistently across species, individuals, and imaging modal- ities [55, 57–59]. While prior work has reported a con- sistent architecture of correlations between regional time series [59], the manner in which the across-region activity pattern at one time point is related to the across-region activity pattern at other time points is not well under- stood. Recent work addressing this gap has posited the existence of so-called lag threads, or spatial similarities between whole-brain activity patterns at non-zero time lags [60]. Unexplained by vasculature, the orthogonal threads are thought to reflect slow, subthreshold changes in the membrane potential of large neuronal populations [61], a notion that is supported by subtle changes in lag thread structure across sleep and during cognitively demanding tasks [60, 62]. Notably, regions of the de- fault mode (which coincide with peaks of the transmodal gradient discussed above) participate in lag thread mo- tifs, where changes in the activity in one region lead to changes in the activity of another region more frequently than expected in an appropriate statistical null model AB β = 0 β = .8 β = 1Trade-o(cid:30) Parameter (β)Relative Temporal Cost (%)Relative Spatial Cost (%) [58, 59]. The existence of these dependencies is consistent with time-invariant, possibly structural features of brain organization producing highly reproducible patterns of activity. RELEVANCE OF NETWORK GEOMETRY FOR DISEASE The spatial architecture of brain networks not only impacts our understanding of dynamics and cognition, but also our understanding of neurological disease and psychiatric disorders. Mounting evidence suggests that many diseases and disorders of mental health can be thought of fruitfully as network disorders, where the anatomy and physiology of cross-regional communication can go awry [63, 64]. Intuitively, spatial anisotropies of developmental processes, spatial specificity of pathology, and spatial inhomogeneities of drug targets that either lead to or accompany these disorders can also explain al- terations in the spatial characteristics of brain networks [65]. In this section we briefly discuss this correspondence in epilepsy, a particularly common neurological disease, and in schizophrenia, a particularly devastating psychi- atric disorder. Despite a diverse pathophysiology and a renitent uni- fying biological mechanism, epilepsy is characterized by altered network dynamics in the form of seizures that display spatially consistent patterns. For example, an ictal period often begins with a marked spatial decor- relation between distributed brain regions followed by a period in which abnormally synchronized activity prop- agates in consistent spatial patterns [66, 67]. In addi- tion to broad patterns of spatial decorrelation, individ- ual siezures also show stereotyped patterns of both spi- ral waves and travelling waves of activity [66, 68–73]. In silico studies have demonstrated that a simple adaptive model of synaptically coupled and spatially embedded excitatory neurons can reproduce many basic features of these waveforms, including their speed and the size of the wavefront [70]. Yet, it is worth noting that travel- ling waves are not unique to epilepsy, but also occur in healthy human and non-human primates where they are thought to play a role in transporting task-relevant infor- mation [71, 74–79]. However, marked differences in wave propagation in healthy and epileptic cortical tissue sug- gests that the precise spatial progression is important, potentially supported by distinct underlying microstruc- tures [80]. Finally, even interictal dynamics are altered in epilepsy, as manifest by marked decreases in average functional connectivity across the brain combined with local increases in functional connectivity and efficiency in default mode areas [81–83]. These connectivity pat- terns have some utility in predicting seizure spread, but the guiding principles leading to these changes and how they relate to fine scale patterns of activity remains un- clear [84]. While its pathophysiology is quite distinct from that 5 implicated in epilepsy, schizophrenia is also a condition marked by severe network disturbances that have broad ramifications for cognitive function [22, 85, 86]. Some of these network alterations appear to selectively affect connections of certain physical lengths, reflecting an al- teration in the network's spatial embedding [87]. Specif- ically, evidence suggests a reduced hierarchical struc- ture and increased connection distance in the anatom- ical connectivity of multimodal cortex in patients with schizophrenia compared to healthy controls, indicative of less efficient spatial wiring [85]. Moreover, in functional brain networks, patients display longer high-weight con- nections, decreased clustering, and increased topological efficiency in comparison to healthy controls [87]. The lack of strong, short distance functional connections is in line with evidence from animal studies suggesting an over-pruning of synapses in childhood onset schizophre- nia [87]. Here, the intuitions gained from a consideration of the network's spatial embedding offer important di- rections for future work in linking non-invasive imaging phenotypes with invasive biomarkers of neural dysfunc- tion in disease. STATISTICS, NULL MODELS, AND GENERATIVE MODELS Collectively, In the previous sections, we outlined developmental rules for efficient wiring and we discussed the reflections of these rules in spatial patterns of healthy and diseased brain dynamics. the studies that we have reviewed motivate the broader use and further development of sophisticated and easily-implementable tools for the analysis of a network's spatial embedding [88]. Here we outline the current state of the field in developing effective network statistics, network null models, and generative network models that account for spatial embedding. Network Statistics. A simple way to examine net- work architecture in the context of spatial embedding is to incorporate the Euclidean distance of connections into local, meso-scale, and global statistics [87, 89]. Ar- guably the simplest local statistics that remain spatially sensitive are moments of the distribution of edge lengths in the network, including the mean, variance, skewness, and kurtosis. One can also compute graph metrics that have been extended to consider space, such as the physi- cal network efficiency and the physical edge betweenness [90]. Intuitively, both begin by defining the length of the shortest physical (as opposed to topological) path along network edges between any two nodes. The physical net- work efficiency then takes the inverse of the harmonic mean of this length, while the physical edge betweenness provides the fraction of shortest physical paths between all node pairs that traverse a given edge [91]. One could also define a physical clustering coefficient in a similar manner. Finally, one can assess the system for Rentian 6 FIG. 2. Spatial distribution of intrinsic neural activity. Principal gradients of functional connectivity calculated in the structural connections of both humans and macaques. The first two principal gradients explained approximately 40% of the observed variance. (A) (Left) A scatter plot of the first two principal gradients, with transmodal regions shown in red, visual regions shown in blue, and sensorimotor regions shown in green. (Right) The same colors are used to show the distribution of points visualized on a cortical surface. The pattern suggests the existence of a macroscale gradient of connectivity that reflects the systematic integration of information across different sensory modalities. (B, Left) The minimum geodesic distance (mm) between each point on the cortical surface and the positive peaks of the first principal gradient. The peaks are shown as white circles. (B, Right) A scatter plot depicting the relationship between distance and location on the trans- to uni-modal gradient. Put differently, transmodal regions with high values in the principal gradient are maximally distant from unimodal regions with low values in the principal gradient. This figure was adapted with permission from [55]. scaling as described earlier, providing information on how efficiently the complex network topology has been embed- ded into the physical space [20, 27–30]. In the context of neural systems, these spatially informed graph statis- tics can be used to account for the physical nature of information processing, propagation, and transmission. Complementing local and global graph statistics is an assessment of a network's community structure, a mesoscale property frequently assessed by considering the existence and strength of network modules [92]. From that community structure, one can determine the spatial embedding of communities, for example by as- sessing their laterality in bilaterally symmetric systems such as the brain [93–95]. One of the most com- mon ways to assess community structure is to maxi- mize a modularity quality function, which identifies as- sortative modules with dense within-module connectivity and sparse between-module connectivity [96] (although see [97] for methods to identify non-assortative commu- nities). Mechanistically, this algorithm compares the strength of observed connections between two nodes in a community to that expected under a given a null model. The most commonly used null model in this context is the Newman-Girvan or configuration model, which pre- serves the strength distribution of the network [96]. How- ever, this null model operationalizes a purely topological constraint – the strength distribution – and does not ac- knowledge any spatial constraints that may exist in the system. For this reason, many investigators across sci- entific domains have begun developing alternative null models that account for physical laws [98] or physical contraints [99–101] on their system of interest. it is worth con- In the context of brain networks, sidering three distinct null models for modularity maximization that incorporate information about the physical space of the network's embedding. First, one can directly incorporate the wiring minimization constraint observed in brain networks by defining a null model with a probability of connection between two nodes that decays exponentially as a function of distance [99]. Using this model, one can detect different, and more spatially distributed modules than those obtained when one uses the configuration model [99]. Second, one can employ gravity models [100], which account for the number of connections expected given a certain distance (typically a power law or inverse of distance), weighted by the relative importance of each location (typically a quantification of the population or size of a given location) [100, 101]. Third, one can employ radiation models designed to capture flow of information between regions, by weighting distance functions by the flux or flow of each location [101]. Of course, there exists no single correct null model for community detection that will suit every question in neuroscience. However, we propose that many studies could test tighter, more targeted hypotheses about community structure in brain networks by using a null model that accounts for the brain's spatial nature. Network Null Models. When considering a network representation of a neural system, one often computes a statistical quantitity of interest and then compares that quantity to that expected in a random network null model. If the observed quantity is significantly greater than or less than that expected, one concludes that the network under study shows meaningful architecture of potential relevance to the biology. Perhaps the most com- mon random network null model is that which randomly permutes the locations of edges in the network while pre- serving the number of nodes, number of edges, and edge 7 ical network are compared to the statistics of each of the generative models with the goal of inferring which wiring rule was most likely to have produced the ob- served architecture [11, 28, 111]. Evidence from such studies suggests that spatially embedded models tend to more accurately reproduce network measures of large- scale neural systems than models that do not account for space [28]. One particularly influential study consid- ered 13 generative models that all incorporated a wiring probability that increased with distance [111]. Consis- tent with other work, the authors found that the model that only included the wiring minimization constraint was unable to recreate long distance connections of in- dividual connectomes in humans [8, 11, 111]. Successive generative models were then added that attempted to recreate certain aspects of topology in addition to these geometric constraints [111]. The models that performed the best were those that preserved homophilic attrac- tion such that connections preferentially formed between nodes that had similar connection profiles [111]. Contin- ued advancement of generative network models, and in- clusion of additional biological features such as bilateral symmetry, serves as an exciting approach to test mech- anistic predictions about how network topology forms in spatially embedded neural systems. FUTURE DIRECTIONS The spatial embedding of the brain is an impor- tant driver of its connectivity, which in turn directly constrains neural function and by extension behavior. Emerging tools from network science can be used to as- sess this spatial architecture, thereby allowing investiga- tors to test more specific hypotheses about brain network structure and dynamics. While we envision that the use of these tools will significantly expand our understand- ing, it is also important to acknowledge their limitations. In particular, the majority of currently available network tools make the simplifying assumption that all of the re- lations of interests are strictly dyadic in nature, and ex- ist between inherently separable components [112]. In truth, however, features that arise from spatial embed- ding can also manifest as continuous or overlapping maps and gradients [53], motivating the use of tools from ap- plied algebraic topology that can account for non-dyadic interactions (Box 2). As the field moves forward, we envision existing and yet-to-be-developed tools for char- acterizing the spatial embedding of brain networks will prove critical for our understanding of network processes underlying cognition, and alterations to those processes accompanying disease. weight distribution. However, one may also be inter- ested to determine whether observed statistics are dif- ferent from what one would expect simply from the spa- tial embedding or wiring rules of the network [102–104]. To address these questions, one can rewire the observed network by conditionally swapping two links if the swap preserves the mean wiring length of the network [103]. By pairing this model with a reduced null model in which connections are only swapped if they reduce connnection length, one can assess the role of long distance connec- tions in the network, which will be preserved in the spa- tial null but not preserved in the reduced null [103]. In addition to preserving the mean wiring length, one might also wish to preserve the full edge length distribution by, for example, (1) fitting a function to the relation- ship between the mean and variance of edge weights and their distances, (2) removing the effect of that relation- ship from the data, (3) randomly rewiring the network, and (4) adding the effect back into the rewired network [102]. To complement insights obtained from edge swapping algorithms, one can also construct null model networks by stipulating a wiring rule a priori while fixing the locations of nodes within the embedded system. In this vein, studies have fruitfully used null models based on minimum spanning tree and greedy triangulation methods [105–107]. A minimum spanning tree is a graph that connects all of the nodes in a network such that the sum of the total edge weights is minimal. To extend this notion to spatial networks, one can preserve the true geographic locations of all nodes in the empirical network and compute the minimum spanning tree on the matrix of Euclidean distances between all node pairs [91]. Representing the opposite extreme is the greedy triangulation model, which is particularly relevant for the study of empirical networks that are planar (lying along a surface) as opposed to non-planar (lying within a volume). In the context of neural systems, planar or planar-like networks are observed in vasculature, and in thinned models of cortex that either consider a single lamina or a coarse-grained model collapsing across laminae [108, 109]. To construct a greedy triangula- tion null model, one can preserve the true geographic locations of all nodes in the empirical network and iteratively connect pairs of nodes in ascending order of their distance while ensuring that no edges cross. After constructing such minimally and maximally wired null models, one can calculate relative measures of wiring length, physical efficiency, physical betweenness centrality, and community structure by normalizing the empirical values by those expected in the two extremes [91]. Generative Network Models. Generative network models can be used to test hypotheses about the rules guiding network growth, development, and evolution [110]. Often, an ensemble of generative models are con- structed, and summary graph statistics from the empir- 8 FIG. 3. Community structure obtained with spatially embedded and non-embedded null models. (A) A schematic of a spatially-informed null model. The model expects fewer long distance connections than short distance connections. (B) A schematic of the anticipated difference between the spatial null model and the Newman-Girvan (NG) null model; spatial communities will have longer distance connections and not capture clustering of spatially nearby regions. (C) Differences in the association matrices between the two models. Positive (negative) numbers indicate when two nodes were more likely to be co-assigned to the same module under the spatial (NG) model. (D) The difference in the participation coefficient between the spatial and NG models. The participation coefficient quantifies how diverse a node's connections are across modules. (E) The difference in spatial spread of modules in both models; the spatially embedded model tends to produce modules that cover larger distances. This figure was adapted with permission from [99]. BOX 1: SIMPLE NETWORK STATISTICS In a network representation of the brain, units rang- ing from neurons or neuronal ensembles to nuclei and areas are represented as network nodes and unit-to-unit interactions ranging from physical connections to statis- tical similarities in activity time series are represented as network edges. The architecture of the network can be quantitatively characterized using statistics from graph theory [113]. Here, we mathematically define some of the topological statistics mentioned elsewhere in this paper. • Degree and Strength. The degree of a node is the number of connections it has. In a binary graph encoded in the adjacency matrix A, where two re- gions i and j are connected if Aij = 1, and not connected if Aij = 0, then the degree ki is defined i,j∈N Aij, where N is the set of all nodes. In a weighted graph, where Aij is the strength of the connection between nodes i and j, then the as ki =(cid:80) strength si is defined as si =(cid:80) i,j∈N Aij. • Path Length and Network Efficiency. The term path length frequently refers to the average length of the shortest path in a network. The short- est path between any two nodes is given by the path requiring the fewest hops. The network ef- ficiency is given by the inverse of the harmonic mean of the shortest path length. To be pre- cise, we can write the path length of node i as Li = 1 , where di,j is the short- n est path length between two nodes and n is the number of nodes. (cid:80) i∈N,j(cid:54)=i di,j (cid:80) n−1 i∈N 2ti i∈N (cid:80) • Clustering Coefficient. The clustering coefficient can be used to quantify the fraction of a node's neighbors that are also neighbors with each other. Specifically, the clustering coefficient of node i is given by Ci = 1 ki(ki−1) , where ti is the n number of triangles around node i [46]. The clus- tering coefficient of the network is the average clus- tering coefficient of all of its nodes. tions exist, the most common is Q = (cid:80) • Modularity. While several modularity quality func- ij[Aij − γPij]δ(cicj), where Q is the modularity quality in- dex, Pij is the expected number of connections be- tween node i and node j under a specified null model, δ() is the Kroenecker delta, and ci indicates the community assignment of node i. The tuning parameter γ ranges from (0,∞) and can be used to tune the average community size. BOX 2: APPLIED ALGEBRAIC TOPOLOGY While graph theory is a powerful and accessible frame- work for analyzing complex networks, complementary in- formation can be gained by using different mathematical 9 formalisms. Here, we describe an alternative approach to studying structure in networks that relies on tools devel- oped in the field of applied algebraic topology, specifically persistent homology [115–117]. Persistent homology can be used to study intrinsically mesoscale structures called cycles and cliques [118]. Cliques are all-to-all connected subsets of nodes in a network. The presence of many, large cliques indicates many highly connected units are present in the network [119]. Cycles are looped patterns of cliques which may enclose a cavity, or topological void, within the network. Cliques and cavities by definition reside within a binary graph, however one can expand a weighted network into a sequence of binary graphs via iterative thresholding [45, 120]. Then using persistent homology one can track the birth, persistence, and death of cavities along this sequence which gives a wholistic insight into the global network (Fig. Box 2, panel A). In random graphs, the number of births and deaths across thresholds follows a characteristic pattern [44]. At high thresholds and low edge density, a few low dimen- sional cavities exist, while at low thresholds and high edge density, more high-dimensional cavities exist (Fig. Box 2, panel B) [44, 121, 122]. Interestingly, geometric graphs – which can be used to instantiate spatial con- straints on the topology – show a markedly different dis- tribution. There are many low dimensional cavities, and fewer cavities with increasing dimension[43, 44, 121, 123] (Fig. Box 2, panel C). This general pattern has been recapitulated in functional networks constructed from firing of hippocampal neurons, indicating a geometric rather than random nature to neuronal co-firing [45]. Furthermore, the persistent homology of human connec- tomes [42] and rat microcircuits [124] is distinct from that expected in a minimally wired null model. In humans, the presence of widespread subcortical connections leads to more cavities being born at high densities [42], while rat microcircuits display more high dimensional cavities in general [124]. Further investigation into how wiring rules shape the topology of neural systems may shed light on how the brain's spatial embedding shapes connectivity across scales and species. BOX 3: CONTROL THEORY Network control theory provides a potentially power- ful approach for modeling neural dynamics [125]. Hail- ing from physics and engineering, network control theory characterizes a complex system as composed of nodes interconnected by edges, and then specifies a model of network dynamics to determine how external input af- fects the nodes' time-varying activity [126]. Most studies of network control in neural systems stipulate a linear, time-invariant model of dynamics: x(t) = Ax(t) + Bu(t), where x is some measure of brain state, A is a struc- tural connectivity matrix, u is the input into the system (exogenous stimulation, or endogenous input from other brain regions), and B selects the control set, or regions 10 FIG. 4. Schematic of network measures. (A) An illustration of network space (topology) and physical space (geometry). The network is embedded into a physical space, indicated by the x- and y-axes. The topological and physical distances between the nodes are not necessarily related. (B) The network representation enables the calculation of local, mesoscale, and global features to describe the pattern of connections in topological space (as shown here) as well as the pattern of connections in physical space (as we describe in the main text). This figure was adapted with permission from [1] and from [114]. to provide input to [127, 128]. Assuming this model of dynamics, one can calculate the control energy required to reach specific brain states, which can be used a state dependent measure of the efficiency of control [129, 130]. Control theory can also posit control metrics that quan- tify how efficiently a node would drive the brain to vari- ous states. Two commonly used metrics are average con- trollability and modal controllability [131]. When every node is included in the control set, average controllabil- ity is proportional to the average energy required to drive the node to any state [132]. Conversely, modal controlla- bility is high in nodes where a small input will result in large perturbations to all eigenmodes of the system, and is interpreted to be high in nodes that can easily drive the brain to hard-to-reach states [133–135]. If these properties are important for helping the brain transition between states, one would expect them not to be randomly distributed across the cortex, but to be clustered into spatially constrained, functionally rele- vant systems. More specifically, one might expect func- tional systems that drive the brain to many accessible states, such as the default mode system, to have high average controllability, while regions that drive the brain to hard-to-reach, cognitively demanding states (executive control areas) to have high modal controllability. Data from healthy human adults supports these two hypothe- ses [133]. Moreover, both average and modal control- lability increase across development and are correlated with cognitive performance generally [134] and executive function specifically [136]. The manner in which network control tracks individual differences reflects the fact that the capacity for a network to enact control is dependent upon its topology [137, 138]. Further efforts are needed to distill exactly how spatial embedding and wiring con- straints impinge on that control capacity, and how it is altered in psychiatric disorders [132] and neurological dis- ease [139]. BOX 4: OUTSTANDING QUESTIONS • How do spatially constrained developmental pro- cesses constrain the formation of cycles in brain networks? • What are the aspects of connectome topology that remain unexplained by wiring minimization or com- munication efficiency and thus may arise from more subtle rules? • How do the structural topologies that arise from physical growth rules support functional gradients? Local Topology (degree, clustering)Path, distancesModule detection, centralityBA 11 • To what extent can we link macro-scale structural topology with small scale developmental rules? • Can a deeper understanding of connectome devel- opment be used to help identify new biomarkers for network diseases? • What additional rules can be incorporated into gen- erative models of the brain to recapitulate its topol- ogy? • How can frontiers in network science help charac- terize non-dyadic relationships in the brain? GLOSSARY • adjacency matrix: The adjacency matrix of a graph is an N × N matrix, where N is the number of nodes. Each element Aij of the matrix gives the strength of the edge between nodes i and j. • cycle: In applied algebraic topology, a cycle is an empty space (or lack of edges) in a graph sur- rounded by all-to-all connected subgraphs of the same dimension. The dimension here refers to the number of nodes included in each all-to-all con- nected subgraph. • edge: From the perspective of graph theory, an edge is a connection between nodes. From the per- spective of neuroscience, an edge is a statistical de- pendency (functional) or estimated physical con- nection (structural) between nodes. • geometry: The geometry of a network reflects fea- tures of a graph in the context of physical space. • hub: A central node in the network, typically hav- ing many connections. • node: From the perspective of graph theory, a node is the unit where edges connect in the graph. From the perspective of neuroscience, a node is a brain region, neuron, or protein whose interactions one wishes to understand. • topology: The quantification of features of a graph in the context of space defined by the graph itself, without respect to any physical embedding. HIGHLIGHTS FIG. 5. Applied algebraic topology. (A) An illustration of thresholding a weighted network across different densities (ρ). At ρ1, a cavity of dimension 1 is born (shown in yellow), which then dies at ρ3. (B) The characteristic pattern of births and deaths (called Betti curves) for cycles of dimension 1 (yellow), dimension 2 (red), and dimension 3 (blue) from a random network. The cliques of each dimension are shown near each corresponding Betti curve for reference. (C) The same pattern, but for geometric networks. Different lines of the same color indicate different dimensions of embedding. This figure was adapted with permission from [45]. • What is the precise relationship between invari- ant features of brain activity and the underlying anatomical structure? • How does the development of the connectome de- termine spatial progression of activity through the cortex in health and disease? • The physical embedding of neural systems imposes constraints on the possible patterns of connections that form, and in turn on the possible repertoire of functional motifs. Utilizing new tools from net- work science that account for these constraints can help researchers understand fundamentals of brain dynamics in health and disease. ABC 12 acknowledge support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the ISI Foundation, the Paul Allen Foundation, the Army Research Laboratory (W911NF-10-2-0022), the Army Research Office (Bassett-W911NF-14-1-0679, Grafton- W911NF-16-1-0474, DCIST-W911NF-17-2-0181), the the National Institute of Office of Naval Research, Mental Health (2-R01-DC-009209-11, R01-MH112847, R01-MH107235, R21-M MH-106799), the National Institute of Child Health and Human Development (1R01HD086888-01), National Institute of Neurological Disorders and Stroke (R01 NS099348), and the National Science Foundation (BCS-1441502, BCS-1430087, NSF PHY-1554488 and BCS-1631550). The content is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies. • Prominent, competing rules guiding the forma- tion of brain networks include the minimization of wiring cost, and the maximization of communica- tion efficiency. These competing mechanisms lead to networks with high local clustering with sparse long distance connections. • Recent work suggests that intrinsic functional con- nectivity varies along dimensions that are tightly linked to the spatial embedding of the brain and the topological properties that arise in the pres- ence of spatial constraints. Similarly, these topo- logical properties show widespread changes in the context of the network diseases such as epilepsy and schizophrenia. • A rich, and continually growing repertoire of statis- tics, null models, and generative models exist to aid researchers in testing focused hypotheses about the role of physical embedding in observed neural phe- nomena. ACKNOWLEDGMENTS We would like to thank Ann Sizemore for her help with Box 2: Applied Algebraic Topology. D.S.B. and J.S. [1] D. S. Bassett and O. Sporns, Nature Neuroscience 20, [15] C. Cherniak, The Journal of Neuroscience 14, 2418 353 (2017), arXiv:0106096v1 [arXiv:cond-mat]. (1994). [2] D. S. Bassett, P. Zurn, and J. I. Gold, Nature Reviews Neuroscience In Press (2018). [3] J. D. Medaglia, M. E. Lynall, and D. S. Bassett, J Cogn Neurosci 27, 1471 (2015). [4] S. E. Petersen and O. Sporns, Neuron 88, 207 (2015). [5] C. Ducruet and L. Beauguitte, Networks and Spatial Economics 14, 297 (2014). [6] M. Barth´elemy, Physics Reports 499, 1 (2011). [7] E. Bullmore and O. Sporns, Nature Reviews Neuro- science 13, 336 (2012), arXiv:NIHMS150003. [8] Y. Chen, S. Wang, C. C. Hilgetag, and C. Zhou, PLoS computational biology 13, e1005776 (2017). [16] C. Cherniak, Z. Mokhtarzada, and R. Rodriguez- Esteban, Evolution of Nervous Systems 1, 269 (2010). [17] A. Raj and Y.-H. Chen, PLoS ONE 6 (2011), 10.1371/. [18] M. Ria Ercsey-Ravasz, N. T. Markov, C. Lamy, D. C. and Van Essen, K. Knoblauch, Z. N. Toroczkai, H. Kennedy, Neuron 80, 184 (2013). [19] H. F. Song, H. Kennedy, and X.-J. Wang, Proceedings of the National Academy of Sciences 111, 16580 (2014). [20] D. S. Bassett, D. L. Greenfield, A. Meyer-Lindenberg, D. R. Weinberger, S. W. Moore, and E. T. Bull- more, PLoS Comput Biol 6 (2010), 10.1371/jour- nal.pcbi.1000748. [9] L. French and P. Pavlidis, PLoS Comput Biol 7 (2011), [21] M. Kaiser and C. C. Hilgetag, PLoS Computational Bi- 10.1371/journal.pcbi.1001049. [10] M. Rubinov, R. J. F. Ypma, C. Watson, and E. T. Bull- more, Proceedings of the National Academy of Sciences 112, 10032 (2015). [11] P. E. Vertes, A. F. Alexander-Bloch, N. Gogtay, J. N. Giedd, J. L. Rapoport, and E. T. Bullmore, Proceed- ings of the National Academy of Sciences 109, 5868 (2012). [12] S. B. Laughlin and T. J. Sejnowski, Science 108, 22 (2002). [13] J. E. Niven and S. B. Laughlin, Journal of Experimental Biology (2008), 10.1242/jeb.017574. [14] J. M. L. Budd and Z. F. Kisv´arday, Frontiers in Neu- roanatomy 6 (2012), 10.3389/fnana.2012.00042. ology 2, 0805 (2006), arXiv:0607034 [q-bio]. [22] A. Zalesky, A. Fornito, G. F. Egan, C. Pantelis, and E. T. Bullmore, Human Brain Mapping 33, 2535 (2012), arXiv:arXiv:1011.1669v3. [23] A. B. Misic, Avena-Koenigsberger, and O. Sporns, Nature Publishing Group 19 (2018), 10.1038/nrn.2017.149. [24] Y. Chen, S. Wang, C. C. Hilgetag, and C. Zhou, PLoS Comput Biol 9 (2013), 10.1371/journal.pcbi.1002937. [25] V. Nicosia, P. E. Vertes, W. R. Schafer, V. Latora, and E. T. Bullmore, Proceedings of the National Academy of Sciences 110, 7880 (2013), arXiv:1304.7364. [26] M. Kaiser, Trends in Cognitive Sciences 21, 703 (2017). [27] M. M. Sperry, Q. K. Telesford, F. Klimm, and D. S. Bassett, Journal of Complex Networks 5, 199 (2017). 13 [28] F. Klimm, D. S. Bassett, J. M. Carlson, P. J. Mucha, and C. C. Hilgetag, PLoS Comput Biol 10 (2014), 10.1371/journal.pcbi.1003491. [29] J. A. Pineda-Pardo, K. Martinez, A. B. Solana, J. A. Hernandez-Tamames, R. Colom, and F. del Pozo, Brain Topogr 28, 187 (2015). [30] A. J. Sadovsky and J. N. MacLean, J Neurosci 34, 7769 (2014). [31] P. Christie and D. Stroobandt, IEEE Trans. on VLSI Systems 8, 639 (2000). [53] S. Jbabdi, S. N. Sotiropoulos, and T. E. Behrens, Current Opinion in Neurobiology 23, 207 (2013), arXiv:NIHMS150003. [54] J. M. Huntenburg, P. L. Bazin, and D. S. Margulies, Trends in Cognitive Sciences 22, 21 (2018). [55] D. S. Margulies, S. S. Ghosh, A. Goulas, M. Falkiewicz, J. M. Huntenburg, G. Langs, G. Bezgin, S. B. Eick- hoff, F. X. Castellanos, M. Petrides, E. Jefferies, and J. Smallwood, Proceedings of the National Academy of Sciences 113, 12574 (2016), arXiv:arXiv:1408.1149. [32] F. Alcalde Cuesta, P. Gonzalez Sequeiros, and [56] M. P. van den Heuvel and O. Sporns, Trends in Cogni- A. Lozano Rojo, Sci Rep 7, 5378 (2017). tive Sciences 17, 683 (2013). [33] R. and D. S. Betzel of F. ceedings ences http://www.pnas.org/content/early/2018/05/07/1720186115.full.pdf. Pro- Sci- 10.1073/pnas.1720186115, the National Academy (2018), Bassett, of [57] H. Lu, Q. Zou, H. Gu, M. E. Raichle, E. A. Stein, and Y. Yang, Proceedings of the National Academy of Sci- ences 109, 3979 (2012). [58] A. Mitra, A. Z. Snyder, T. Blazey, and M. E. Raichle, [34] J. A. Henderson and P. A. Robinson, Physical Review Letters 107 (2011), 10.1103/PhysRevLett.107.018102. [35] H. J. Park and K. Friston, Science 342 (2013), 10.1126/science.1238411. [36] O. Sporns, D. R. Chialvo, M. Kaiser, and C. C. Hilge- tag, Trends in Cognitive Sciences 8, 418 (2004). [37] M. P. van den Heuvel, R. S. Kahn, J. Goni, and O. Sporns, Proceedings of the National Academy of Sci- ences 109, 11372 (2012). [38] J. Seidlitz, F. V´asa, M. Shinn, R. Romero-Garcia, K. J. Whitaker, P. E. V´ertes, K. Wagstyl, P. Kirk- patrick Reardon, L. Clasen, S. Liu, A. Messinger, D. A. Leopold, P. Fonagy, R. J. Dolan, P. B. Jones, I. M. Goodyer, A. Raznahan, and E. T. Bullmore, Neuron 97, 231 (2018). [39] S. Varier and M. Kaiser, PLoS Comput Biol 7 (2011), 10.1371/journal.pcbi.1001044. [40] C. C. Hilgetag and A. Goulas, Brain Struct Funct 221, 2361 (2016). [41] R. Perin, T. K. Berger, and H. Markram, Proceedings of the National Academy of Sciences 108, 5419 (2011), arXiv:arXiv:1408.1149. [42] A. E. Sizemore, C. Giusti, A. Kahn, J. M. Vettel, R. F. Betzel, and D. S. Bassett, Journal of Computational Neuroscience 44, 1 (2017), arXiv:1608.03520. [43] M. Kahle, "Random Geometric Complexes," (2018). [44] M. Kahle, Discrete Mathematics 309, 1658 (2009), arXiv:0605536 [math]. [45] C. Giusti, E. Pastalkova, C. Curto, and V. It- skov, Proc Natl Acad Sci U S A 112, 13455 (2015), arXiv:1502.06172. [46] D. J. Watts and S. H. Strogatz, Nature 393, 440 (1998), arXiv:0803.0939v1. [47] C. T. Shih, O. Sporns, S. L. Yuan, T. S. Su, Y. J. Lin, C. C. Chuang, T. Y. Wang, C. C. Lo, R. J. Greenspan, and A. S. Chiang, Curr Biol 25, 1249 (2015). [48] M. Kaiser and C. C. Hilgetag, Phys Rev E Stat Nonlin Soft Matter Phys 69, 036103 (2004). Proc Natl Acad Sci 17112, 2235 (2015). [59] M. E. Raichle, Annual Review of Neuroscience 38, 433 (2015), arXiv:0402594v3 [arXiv:cond-mat]. [60] A. Mitra, A. Z. Snyder, C. D. Hacker, and M. E. Raichle, Journal of Neurophysiology 111, 2374 (2014). [61] A. Mitra, A. Kraft, P. Wright, B. Acland, A. Z. Sny- der, Z. Rosenthal, L. Czerniewski, A. Bauer, L. Snyder, J. Culver, J.-M. Lee, and M. E. Raichle, Neuron (2018), https://doi.org/10.1016/j.neuron.2018.03.015. [62] A. Mitra, A. Z. Snyder, E. Tagliazucchi, H. Laufs, and M. E. Raichle, eLife 4, 1 (2015). [63] U. Braun, A. Schaefer, R. F. Betzel, H. Tost, A. Meyer- Lindenberg, and D. S. Bassett, Neuron 97, 14 (2018). [64] C. J. Stam, Nat Rev Neurosci 15, 683 (2014). [65] D. S. Bassett, C. H. Xia, and T. D. Satterthwaite, Biol Psychiatry Cogn Neurosci Neuroimaging S2451-9022, 30079 (2018). [66] V. K. Jirsa, W. C. Stacey, P. P. Quilichini, A. I. and C. Bernard, Brain 137, 2210 (2014), Ivanov, arXiv:arXiv:1011.1669v3. [67] F. Wendling, F. Bartolomei, J. J. Bellanger, J. Bourien, and P. Chauvel, Brain 126, 1449 (2003). [68] A. C. Chamberlain, J. Viventi, J. A. Blanco, D. H. Kim, J. A. Rogers, and B. Litt, Proceedings of the Annual International Conference of the IEEE Engineering in Medicine and Biology Society, EMBS , 761 (2011). [69] J. Viventi, D. H. Kim, L. Vigeland, E. S. Frechette, J. A. Blanco, Y. S. Kim, A. E. Avrin, V. R. Tiruvadi, S. W. Hwang, A. C. Vanleer, D. F. Wulsin, K. Davis, C. E. Gelber, L. Palmer, J. Van Der Spiegel, J. Wu, J. Xiao, Y. Huang, D. Contreras, J. A. Rogers, and B. Litt, Nature Neuroscience 14, 1599 (2011). [70] L. R. Gonz´alez-Ram´ırez, O. J. Ahmed, S. S. Cash, C. E. Wayne, and M. A. Kramer, PLoS Computational Biol- ogy 11 (2015), 10.1371/journal.pcbi.1004065. [71] K. A. Richardson, S. J. Schiff, and B. J. Gluck- man, Physical Review Letters 94 (2005), 10.1103/Phys- RevLett.94.028103. [49] D. S. Bassett and E. T. Bullmore, Neuroscientist 23, [72] M. Ursino and G. E. La Cara, Journal of Theoretical 499 (2017), arXiv:1608.05665. Biology 242, 171 (2006). [50] A. Avena-Koenigsberger, J. Goni, R. Sole, and O. Sporns, Journal of The Royal Society Interface 12, 20140881 (2014). [51] L. H. Scholtens, R. Schmidt, M. A. de Reus, and M. P. van den Heuvel, J Neurosci 34, 12192 (2014). [52] M. P. van den Heuvel, L. H. Scholtens, L. Feldman Bar- rett, C. C. Hilgetag, and M. A. de Reus, J Neurosci 35, 13943 (2015). [73] L. E. Martinet, G. Fiddyment, J. R. Madsen, E. N. Eskandar, W. Truccolo, U. T. Eden, S. S. Cash, and M. A. Kramer, Nature Communications 8 (2017), 10.1038/ncomms14896. [74] J. M. Beggs and N. Timme, Frontiers in Physiology 3 JUN, 1 (2012). [75] M. Massimini, Journal of Neuroscience 24, 6862 (2004). [76] D. Rubino, K. A. Robbins, and N. G. Hatsopoulos, Nature Neuroscience 9, 1549 (2006). [77] K. Takahashi, S. Kim, T. P. Coleman, K. A. Brown, A. J. Suminski, M. D. Best, and N. G. Hatsopoulos, Na- ture Communications 6 (2015), 10.1038/ncomms8169. [78] K. Takahashi, M. Saleh, R. D. Penn, and N. G. Hat- sopoulos, Frontiers Human Neuroscience 5, 40 (2011). [79] H. Zhang, A. J. Watrous, A. Patel, and J. Jacobs, Neu- ron 98, 1269 (2018). [80] A. Benucci, R. A. Frazor, and M. Carandini, Neuron 55, 103 (2007). [81] L. Douw, M. N. DeSalvo, N. Tanaka, A. J. Cole, H. Liu, C. Reinsberger, and S. M. Stufflebeam, Annals of Clin- ical and Translational Neurology 2, 338 (2015). [82] L. Bonilha, T. Nesland, G. U. Martz, J. E. Joseph, M. V. Spampinato, J. C. Edwards, and A. Tabesh, J Neurol Neurosurg Psychiatry 83, 903 (2012). [83] M. N. DeSalvo, L. Douw, N. Tanaka, C. Reinsberger, and S. M. Stufflebeam, Radiology (2014), 10.1148/ra- diol.13131044. [84] V. K. Jirsa, T. Proix, D. Perdikis, M. M. Woodman, H. Wang, C. Bernard, C. Bınar, P. Chauvel, F. Bar- tolomei, F. Bartolomei, M. Guye, J. Gonzalez-Martinez, and P. Chauvel, NeuroImage 145, 377 (2017). [85] D. S. Bassett, E. Bullmore, B. A. Verchinski, V. S. Mat- tay, D. R. Weinberger, and A. Meyer-Lindenberg, Jour- nal of Neuroscience 28, 9239 (2008). [86] A. Griffa, P. S. Baumann, J. P. Thiran, and P. Hag- mann, NeuroImage 80, 515 (2013). [87] A. F. Alexander-Bloch, P. E. V´ertes, R. Stidd, F. Lalonde, L. Clasen, J. Rapoport, J. Giedd, E. T. Bull- more, and N. Gogtay, Cerebral Cortex 23, 127 (2013). [88] M. Kaiser, NeuroImage 57, 892 (2011). [89] J. M. Duarte-Carvajalino, N. Jahanshad, C. Lenglet, K. L. McMahon, G. I. De Zubicaray, N. G. Martin, M. J. Wright, P. M. Thompson, and G. Sapiro, NeuroImage 59, 3784 (2012), arXiv:NIHMS150003. [90] J. Buhl, J. Gautrais, R. V. Sole, P. Kuntz, S. Valverde, J. L. Deneubourg, and G. Theraulaz, The European Physical Journal B 42, 123 (2004). [91] L. Papadopoulos, P. Blinder, H. Ronellenfitsch, F. Klimm, E. Katifori, D. Kleinfeld, and D. S. Bas- sett, arxiv 1612, 08058 (2018). 14 [100] P. Expert, T. Evans, V. D. Blondel, and R. Lam- biotte, Proceedings of the National Academy of Sciences (2010), 10.1073/pnas.1018962108, arXiv:1012.3409. [101] M. Sarzynska, E. A. Leicht, G. Chowell, M. A. Porter, D. Bassett, and M. Sarzynska, Journal of Complex Net- works (2015), 10.1093/comnet/cnv027. [102] J. A. Roberts, A. Perry, A. R. Lord, G. Roberts, P. B. Mitchell, R. E. Smith, F. Calamante, and M. Breaks- pear, NeuroImage 124, 379 (2016). [103] D. Samu, A. K. Seth, and T. Nowotny, PLoS Computa- tional Biology 10 (2014), 10.1371/journal.pcbi.1003557. [104] M. Wiedermann, J. F. Donges, J. Kurths, and R. V. Donner, Physical Review E 93 (2016), 10.1103/Phys- RevE.93.042308, arXiv:1509.09293. [105] X. Cui, J. Xiang, H. Guo, G. Yin, H. Zhang, F. Lan, and J. Chen, Front Comput Neurosci 12, 31 (2018). [106] T. W. P. Janssen, A. Hillebrand, A. Gouw, K. Gelade, R. Van Mourik, A. Maras, and J. Oosterlaan, Clin Neu- rophysiol 128, 2258 (2017). [107] D. J. Smit, E. J. de Geus, M. Boersma, D. I. Boomsma, and C. J. Stam, Brain Connect 6, 312 (2016). [108] P. Blinder, P. S. Tsai, J. P. Kaufhold, P. M. Knutsen, H. Suhl, and D. Kleinfeld, Nat Neurosci 16, 889 (2013). and [109] F. Schmid, P. S. Tsai, D. Kleinfeld, P. Jenny, B. Weber, PLoS Comput Biol 13, e1005392 (2017). [110] R. F. Betzel and D. S. Bassett, J R Soc Interface 14, 20170623 (2017). [111] R. F. Betzel, A. Avena-Koenigsberger, J. Goni, Y. He, M. A. de Reus, A. Griffa, P. E. V´ertes, B. Misic, J. P. Thiran, P. Hagmann, M. van den Heuvel, X. N. Zuo, E. T. Bullmore, and O. Sporns, NeuroImage 124, 1054 (2016), arXiv:1506.06795. [112] C. T. Butts, Science 325, 414 (2009), arXiv:arXiv:1010.0725v1. [113] M. Rubinov and O. Sporns, NeuroImage 52, 1059 (2010). [114] J. O. Garcia, A. Ashourvan, S. F. Muldoon, J. M. Vettel, and D. S. Bassett, Proceedings of the IEEE 106, 846 (2018). [115] C. Giusti, R. Ghrist, Neurosci 41, 1 (2016). and D. S. Bassett, J Comput [116] A. Zomorodian and G. Carlsson, Discrete and Compu- tational Geometry 33, 249 (2005). [92] S. Fortunato and D. Hric, Physics Reports 659, 1 [117] G. Carlsson, Bulletin of the American Mathematical So- (2016). [93] K. W. Doron, D. S. Bassett, and M. S. Gazzaniga, Proceedings of the National Academy of Sciences 109, 18661 (2012). [94] L. R. Chai, M. G. Mattar, I. A. Blank, E. Fedorenko, and D. S. Bassett, Cereb Cortex 26, 4148 (2016). [95] X. He, D. S. Bassett, G. Chaitanya, M. R. Sperling, ciety 46, 255 (2009), arXiv:arXiv:1312.6184v5. [118] A. E. Sizemore, J. Phillips-Cremins, R. Ghrist, and D. S. Bassett, arXiv 1806, 05167 (2018). [119] P. Dotko, K. Hess, R. Levi, M. Nolte, M. Reimann, M. Scolamiero, K. Turner, E. Muller, and H. Markram, Frontiers in Computational Neuroscience 11, 48 (2016), arXiv:1601.01580. L. Kozlowski, and J. I. Tracy, Brain , 141 (2018). [120] G. Petri, M. Scolamiero, I. Donato, F. Vaccarino, and [96] M. E. J. Newman and M. Girvan, Physical Review E (2003), 10.1103/PhysRevE.69.026113, arXiv:0308217 [cond-mat]. R. Lambiotte, PLoS ONE 8 (2013), 10.1371/. [121] A. Sizemore, C. Giusti, and D. S. Bassett, Journal of Complex Networks 5, 245 (2017), arXiv:1512.06457. [97] R. F. Betzel, J. D. Medaglia, and D. S. Bassett, Nat [122] D. Horak, S. Maleti, and M. Rajkovi, Stat. Mech Commun 9, 346 (2018). [98] L. Papadopoulos, J. G. Puckett, K. E. Daniels, and D. S. Bassett, Phys Rev E 94, 032908 (2016). [99] R. F. Betzel, J. D. Medaglia, L. Papadopoulos, G. Baum, R. Gur, R. Gur, D. Roalf, T. D. Satterth- waite, and D. S. Bassett, Network Neuroscience (2016), arXiv:1608.01161. (2009), 10.1088/1742-5468/2009/03/P03034. [123] O. Bobrowski and M. Kahle, Journal of Applied and Computational Topology 1, 10.1007/s41468-017-0010-0. [124] P. Dotko, K. Hess, R. Levi, M. Nolte, M. Reimann, M. Scolamiero, K. Turner, E. Muller, and H. Markram, Frontiers in Computational Neuroscience 1, 1 (2016), 1601.01580. [125] S. J. Schiff, Neural Control Enginerring. The emerg- ing intersection between control theory and neuroscience (MIT Press, 2011). J. [126] Y. Y. and A. Liu, J. Slotine, 473, L. (2011), 15 [133] S. Gu, F. Pasqualetti, M. Cieslak, Q. K. Telesford, A. B. Yu, A. E. Kahn, J. D. Medaglia, J. M. Vettel, M. B. Miller, S. T. Grafton, and D. S. Bassett, Nature Com- munications 6, 1 (2015), arXiv:1406.5197. [134] E. Tang, C. Giusti, G. L. Baum, S. Gu, E. Pollock, A. E. Kahn, D. R. Roalf, T. M. Moore, K. Ruparel, R. C. Gur, R. E. Gur, T. D. Satterthwaite, and D. S. Bassett, Nature Communications 8 (2017), 10.1038/s41467-017- 01254-4, arXiv:1607.01010. [135] E. Wu-Yan, R. F. Betzel, E. Tang, S. Gu, F. Pasqualetti, and D. S. Bassett, Journal of Nonlinear Science (2017), arXiv:1706.05117. [136] E. J. Cornblath, E. Tang, G. L. Baum, T. M. Moore, D. R. Roalf, R. C. Gur, R. E. Gur, F. Pasqualetti, T. D. Satterthwaite, and D. S. Bassett, arXiv 1801, 04623 (2018). [137] T. Menara, V. Katewa, D. S. Bassett, and Nature Barab´asi, arXiv:/www.nature.com/nature/journal/v473/n7346/abs/10.1038- nature10011-unlocked.html#supplementary- information [http:]. 167 [127] A. E. Bryson, IEEE Control Systems 16, 26 (1996). [128] G. Yan, P. E. Vertes, E. K. Towlson, Y. L. Chew, D. S. Walker, W. R. Schafer, and A. L. Barabasi, Nature 550, 519 (2017). [129] S. Gu, R. F. Betzel, M. G. Mattar, M. Cieslak, P. R. Delio, S. T. Grafton, F. Pasqualetti, and D. S. Bassett, Neuroimage 148, 305 (2017). [130] R. F. Betzel, S. Gu, J. D. Medaglia, F. Pasqualetti, and D. S. Bassett, Sci Rep 6, 30770 (2016). [131] F. Pasqualetti, S. Zampieri, and F. Bullo, IEEE Trans- F. Pasqualetti, arXiv 1706, 05120 (2018). actions on Control of Network Systems 1, 40 (2014). [132] J. Jeganathan, A. Perry, D. S. Bassett, G. Roberts, P. B. Mitchell, and M. Breakspear, NeuroImage Clinical In Press (2018). [138] J. Z. Kim, J. M. Soffer, A. E. Kahn, J. M. Vettel, F. Pasqualetti, and D. S. Bassett, Nat Phys 14, 91 (2018). [139] P. N. Taylor, J. Thomas, N. Sinha, J. Dauwels, M. Kaiser, T. Thesen, and J. Ruths, Front Neurosci 9, 202 (2015).
1803.03304
2
1803
2019-03-01T23:25:16
A model of reward-modulated motor learning with parallelcortical and basal ganglia pathways
[ "q-bio.NC", "cs.NE" ]
Many recent studies of the motor system are divided into two distinct approaches: Those that investigate how motor responses are encoded in cortical neurons' firing rate dynamics and those that study the learning rules by which mammals and songbirds develop reliable motor responses. Computationally, the first approach is encapsulated by reservoir computing models, which can learn intricate motor tasks and produce internal dynamics strikingly similar to those of motor cortical neurons, but rely on biologically unrealistic learning rules. The more realistic learning rules developed by the second approach are often derived for simplified, discrete tasks in contrast to the intricate dynamics that characterize real motor responses. We bridge these two approaches to develop a biologically realistic learning rule for reservoir computing. Our algorithm learns simulated motor tasks on which previous reservoir computing algorithms fail, and reproduces experimental findings including those that relate motor learning to Parkinson's disease and its treatment.
q-bio.NC
q-bio
A model of reward-modulated motor learning with parallel cortical and basal ganglia pathways Ryan Pyle1,∗ and Robert Rosenbaum1,2,∗ 1: Department of Applied and Computational Mathematics and Statistics, University of Notre Dame, Notre Dame, Indiana, USA. 2:Interdisciplinary Center for Network Science and Applications, University of Notre Dame, Notre Dame, Indiana, USA. * [email protected], [email protected] Acknowledgments This work was supported by National Science Foundation grants DMS-1517828, DMS-1654268, and DBI-1707400. The authors thank Jonathan Rubin, Robert Turner, and Robert Mendez for helpful discussions. Abstract Reservoir computing is a biologically inspired class of learning algorithms in which the intrinsic dynamics of a recurrent neural network are mined to produce target time series. Most existing reservoir computing algorithms rely on fully supervised learning rules, which require access to an exact copy of the target response, greatly reducing the utility of the system. Reinforcement learning rules have been developed for reservoir computing, but we find that they fail to converge on complex motor tasks. Current theories of biological motor learning pose that early learning is controlled by dopamine modulated plasticity in the basal ganglia that trains parallel cortical pathways through unsupervised plasticity as a motor task becomes well-learned. We developed a novel learning algo- rithm for reservoir computing that models the interaction between reinforcement and unsupervised learning observed in experiments. This novel learning algorithm converges on simulated motor tasks on which previous reservoir computing algorithms fail, and reproduces experimental findings that relate Parkinson's disease and its treatments to motor learning. Hence, incorporating biological theories of motor learning improves the effectiveness and biological relevance of reservoir computing models. 9 1 0 2 r a M 1 ] . C N o i b - q [ 2 v 4 0 3 3 0 . 3 0 8 1 : v i X r a 1 Introduction Even simple motor tasks require intricate, dynamical patterns of muscle activations. Understanding how the brain generates this intricate motor output is a central problem in neuroscience that can inform the development of brain-machine interfaces, treatments for motor diseases, and control algorithms for robotics. Recent work is largely divided into addressing two distinct questions: How are motor responses encoded and how are they learned? From the coding perspective, it has been shown that the firing rates of cortical neurons ex- hibit intricate dynamics that do not always code for specific stimulus or movement parame- ters (Churchland et al., 2012; Russo et al., 2018). A prevailing theory poses that these firing rate patterns are part of an underlying dynamical system that serves as a high-dimensional "reservoir" of dynamics from which motor output signals are distilled (Shenoy, Sahani, & Churchland, 2013; Sussillo, 2014). This notion can be formalized by reservoir computing models, in which a chaotic or near chaotic, recurrent neural network serves as a reservoir of firing rate dynamics and synaptic readout weights are trained to produce target time series (Maass, Natschlager, & Markram, 2002; Jaeger & Haas, 2004; Sussillo & Abbott, 2009; Lukosevicius, Jaeger, & Schrauwen, 2012; Sussillo, 2014). Reservoir computing models can learn to generate intricate dynamical responses and natu- rally produce firing rate dynamics that are strikingly similar to those of cortical neurons (Sussillo, Churchland, Kaufman, & Shenoy, 2013; Mante, Sussillo, Shenoy, & Newsome, 2013; Laje & Buono- mano, 2013; Hennequin, Vogels, & Gerstner, 2014). However, most reservoir computing models rely on biologically unrealistic, fully supervised learning rules. Specifically, they must learn from a teacher signal that can already generate the target output. Many motor tasks are not learned in an environment in which such a teacher signal is available. Instead, motor learning is at least partly realized through reward-modulated, reinforcement learning rules (Izawa & Shadmehr, 2011). A large body of studies are committed to understanding how reinforcement learning is im- plemented in the motor systems of mammals and songbirds (Brainard & Doupe, 2002; Olveczky, Andalman, & Fee, 2005; Kao, Doupe, & Brainard, 2005; Ashby, Turner, & Horvitz, 2010; Izawa & Shadmehr, 2011; Fee, 2014). The basal ganglia and their homolog in songbirds play a critical role in reinforcement learning of motor tasks through dopamine-modulated plasticity at corticostriatal synapses. This notion inspired the development of a reward-modulated learning rule for reservoir computing (Hoerzer, Legenstein, & Maass, 2014). However, we found that this learning rule fails to converge on many simulated motor tasks. We propose that the shortcomings of previous reservoir computing models can be resolved by a closer inspection of the literature on biological motor learning. A large body of evidence across multiple species supports a theory of learning in which dopamine-modulated plasticity in the basal ganglia or its homologs is responsible for early learning and this pathway gradually trains a parallel cortical pathway that takes over as tasks become well learned or "automatized" (Bottjer, Miesner, & Arnold, 1984; Carelli, Wolske, & West, 1997; Brainard & Doupe, 2000; Pasupathy & Miller, 2005; Ashby, Ennis, & Spiering, 2007; Obeso et al., 2009; Andalman & Fee, 2009; Ashby et al., 2010; Turner & Desmurget, 2010; Fee & Goldberg, 2011; Olveczky, Otchy, Goldberg, Aronov, & Fee, 2011), although the biology is not settled (Kawai et al., 2015). This model of motor learning has only been tested computationally in discrete choice tasks that do not capture the intricate, dynamical nature of motor responses (Ashby et al., 2007). Inspired by this theory of automaticity from parallel pathways, we derived a new architecture 2 Figure 1: Network diagrams for three reservoir computing algorithms. A) In FORCE learning, readout weights are trained to match a target using a fully supervised error signal. Output is fed back to the reservoir. B) RMHL is similar to FORCE, but learning is driven by a reward- modulated plasticity and exploratory noise. C) SUPERTREX combines elements of FORCE and RMHL. The exploratory pathway (green) is driven by noise, and acts similarly to RMHL. The mastery pathway (purple) is analogous to FORCE, but uses the output of the exploratory pathway in place of the fully supervised signal used by FORCE. The sum of both pathways provides the total output. and learning rule for reservoir computing. In this model, a reward-modulated pathway is respon- sible for early learning and serves as a teacher signal for a parallel pathway that takes over the production of motor output as the task becomes well-learned. This algorithm is applicable to a large class of motor learning tasks to which fully supervised learning models cannot be applied and it outperforms previous reward-modulated models. We additionally show that our model naturally produces experimental and clinical findings that relate Parkinson's disease and its treatment to motor learning (Ashby et al., 2007, 2010; Turner & Desmurget, 2010). Results We first review two previous learning rules for reservoir computing, then introduce a new, biologi- cally inspired learning rule that combines their strengths. FORCE learning One of the most powerful and widely used reservoir computing algorithms is first-order reduced and controlled error or FORCE (Sussillo & Abbott, 2009), which is able to rapidly and accurately learn to generate complex, dynamical outputs. The standard architecture for FORCE is schematized in Fig. 1A (FORCE variants exist, although the underlying principle is the same). The reservoir is composed of a recurrently connected population of "rate-model" neurons. The output of the reservoir is trained to produce a target time series by modifying a set of readout weights, and the output affects the reservoir through a feedback loop. The reservoir dynamics are defined by = −x + Jr + Qz τ dx dt 3 (1) outputrecurrent netFORCESUPERTREXReward-modulated plasticityUnsupervised plasticityFully supervised plasticityrecurrent netRMHLexploratorynoiserecurrent netoutputoutputexploratorynoiseABC Here, r = tanh(x) +  is a time-dependent vector representing the activity of units within the reservoir, τ is a time constant,  is a small noise term, z = W r is the reservoir output, J is the recurrent connectivity matrix, Q the feedback weights and W the readout weights. A feedforward input term is also commonly included (Sussillo & Abbott, 2009), but we do not use one here. When the spectral radius of J is sufficiently large, the intrinsic dynamics of r(t) become rich and chaotic (Sompolinsky, Crisanti, & Sommers, 1988). The goal of FORCE is to utilize these rich dynamics by modifying readout weights, W , in such a way that the output, z(t), matches a desired target function, f (t). A powerful and widely used learning algorithm for FORCE, recursive least squares (RLS), is defined by (Sussillo & Abbott, 2009) where τw dW dt = −erT P e(t) = z(t) − f(t) (2) (3) is the error vector and τw is the learning time scale. The matrix, P , is a running estimate of the inverse of the correlation matrix of rates, r (see Materials and Methods). FORCE excels at generating a target time series by harvesting reservoir dynamics, but is incom- plete as a model of motor learning. As a fully supervised learning rule, FORCE must have access to the correct output to determine its error (see the presence of f in Eq. (3)). Since the correct output must already be generated to compute the error, FORCE can only learn target functions that are already known explicitly and can be generated. Many motor learning tasks require the generation of an unknown target using a lower-dimensional error signal (Izawa & Shadmehr, 2011). We consider examples of such tasks below. A potential solution for these issues is provided by appealing to biological motor learning, which is controlled at least in part by dopamine-modulated reinforcement learning in the basal ganglia (Turner & Desmurget, 2010; Ashby et al., 2010; Izawa & Shadmehr, 2011). Reward modulated Hebbian learning Reward-modulated Hebbian learning (RMHL) (Hoerzer et al., 2014) is a reinforcement learning rule for reservoir computing in which reward is indicated by a one-dimensional error signal using a plasticity rule inspired by dopamine-dependent Hebbian plasticity observed in the basal ganglia. RMHL uses the same reservoir dynamics (Eq. (1)) and the same basic architecture as FORCE (Fig. 1B), but the learning rule is fundamentally different. The original RMHL algorithm (Hoerzer et al., 2014) used a binary error signal, despite poten- tially poorer performance than other options, to demonstrate that the algorithm could learn with minimal information. We implement a modified version of RMHL with an error signal, e(z(t), t), that can be any time-dependent, non-negative function of the output, z, which the algorithm will seek to minimize. Equivalently, e is proportional to the negative of reward. 4 In contrast to the fully supervised error signal, e, used in FORCE learning, e is scalar (one- dimensional) even when z is a higher-dimensional vector. Moreover, e can quantify any notion of "error" or "cost" associated with the output, z, and does not assume that a target output is known or even that there exists a unique target output. This allows RMHL to be applied to a large class of learning tasks to which FORCE cannot be applied, as we demonstrate below. To decrease error, RMHL makes random perturbations to the reservoir output as a form of exploration. Specifically, the output, z, is given by z = W r + Ψ(e)η where η(t) is exploratory noise and Ψ is a sublinear function that serves to damp runaway oscilla- tions during learning. The learning rule is then given by τw dW dt = Φ(e)zrT (4) where x denotes a high-pass filtered version of x, which represents recent changes in x and Φ is a sublinear function that controls when to update the weights. We assume that Ψ is an increasing function and Φ an odd function with Ψ(0) = Φ(0) = 0. This assures that exploration and learning are effectively quenched when the error is consistently near zero. Intuitively, the learning rule can be understood as follows: If a random perturbation from η has recently decreased e, this perturbation is then incorporated into W . SUPERTREX: A new learning algorithm for reservoir computing Unfortunately, on many tasks, the weights trained by RMHL fail to converge to an accurate solution, as we show below. RMHL models dopamine-modulated learning in the basal ganglia, but does not account for experimental evidence for the eventual independence of well-learned tasks on the activity of the basal ganglia: It has been proposed that the basal ganglia are responsible for early learning, but train a parallel cortical pathway that gradually takes over the generation of output as tasks become well-learned and "automatized" (Pasupathy & Miller, 2005; Ashby et al., 2007, 2010; Turner & Desmurget, 2010; H´elie, Paul, & Ashby, 2012). This could explain why some neurons in the basal ganglia are active during early learning and exploration, but inactive as the task becomes well-learned (Carelli et al., 1997; Miyachi, Hikosaka, & Lu, 2002; Pasupathy & Miller, 2005; Poldrack et al., 2005; Ashby et al., 2007; Tang et al., 2009; Ashby et al., 2010; H´elie et al., 2012). It could also explain why patients or animals with basal ganglia lesions can perform previously learned tasks well, but suffer impairments at learning new tasks (Miyachi, Hikosaka, Miyashita, K´ar´adi, & Rand, 1997; Obeso et al., 2009; Turner & Desmurget, 2010). This idea is also consistent with many findings suggesting that the basal ganglia homolog in song birds is responsible for early learning and exploration of novel song production, but not for the vocalization of well-learned songs (M. S. Brainard, 2004; Kao et al., 2005; Aronov, Andalman, & Fee, 2008; Andalman & Fee, 2009; Fee & Goldberg, 2011). The FORCE and RMHL algorithms could be seen as analogous to the individual pathways in this theory of motor learning: RMHL learns through reward-modulated exploration analogous to the basal ganglia, while FORCE models cortical pathways that learn from the output produced by the basal ganglia pathway. Inspired by this analogy, we next introduce a new algorithm, Supervised 5 Learning Trained by Rewarded Exploration (SUPERTREX), that combines the strengths RMHL and FORCE to overcome the limitations of each. The architecture of SUPERTREX (Fig. 1C) is different than the architectures of FORCE and RMHL: There are now two distinct sets of weights from the reservoir to the outputs, and each is trained with a separate learning rule. The "exploratory pathway" learns via an RMHL-like, reinforcement learning algorithm, requiring only a one-dimensional metric of performance rather than an explicit error signal. The exploratory pathway is roughly based off of the biological basal ganglia pathway. The "mastery pathway" learns through a FORCE-like algorithm. The key idea is that the activity of the exploratory pathway can act as a target for the mastery pathway to learn, replacing the supervised error signal required by FORCE. Hence, SUPERTREX does not need the explicit supervisory error signal that FORCE does. The mastery pathway is roughly based off of the biological cortical pathway. Importantly, the convergence issues we have found with RMHL are not problematic for SU- PERTREX because weights in the RMHL-like exploratory pathway do not need to converge to a correct solution because weights in the mastery pathway converge instead. SUPERTREX uses the same reservoir dynamics (Eqs. (1)), but the outputs are determined by z1 = W1r + Ψ(e)η z2 = W2r z = z1 + z2. Here, z1 is the output from the exploratory pathway, z2 from the mastery pathway, and z is the total output. The learning rules are defined by τw1 τw2 dW1 dt dW2 dt = Φ(e)z rT = (z − z2)rT P. (5) Intuitively, the first learning rule works like RMHL to quickly minimize the total error, as it uses the error of the total output, z. However, it only controls the z1 component of z, resulting in z1 + z2 ≈ f. Error between the z2 component and f is therefore just z1, which replaces the error in the second learning rule since z− z2 = z1. As z2 approaches f, learning in the exploratory pathway causes z1 to approach 0 in order to keep the total z correct. Additionally, we added one extra component to the SUPERTREX algorithm. Learning transfer from the exploratory pathway is soft thresholded based on total error - if error grows above this point, the transfer rate is gradually reduced to 0. This means that transfer can only occur if the total combined output of both pathways is correct. In practice, this is true for the entire learning period except for a small initial period while the exploratory pathway is finding a solution. Performance without this addition was similar overall, but slightly slower. Note that the learning rule for W1 is local in the sense that it only involves values of the presynaptic and postsynaptic variables in addition to the error signal, e. The learning rule for W2 would be local if it were not for the computation of P , which is biologically unrealistic. However, P can be replaced by the identity matrix to make the learning rules for SUPERTREX purely local. This slows down learning, but the network can still learn to produce target outputs from a one-dimensional error signal (see (Hoerzer et al., 2014) and the disrupted learning example below). 6 In summary, RMHL-like learning in the exploratory pathway uses a one-dimensional error signal, e, to track the target while FORCE-like learning in the mastery pathway uses the exploratory pathway as a teacher signal until it learns the output and takes over. This models current theories of biological motor learning in which early learning is dominated by dopamine-dependent plasticity in the basal ganglia, which gradually trains parallel cortical pathways as the task becomes well- learned. We next test SUPERTREX on three increasingly difficult motor tasks, comparing the perfor- mance of SUPERTREX to that of FORCE and RMHL. Task 1: Generating a known target output We first consider a task in which the goal is to draw a parameterized curve of a butterfly by directly controlling the coordinates of a pen (Fig. 2A). Specifically, the target is given by f (t) = (x(t), y(t)) where x(t) and y(t) parameterize the x- and y-coordinates of a pen that successfully traces out the butterfly. The reservoir output, z(t), controls the coordinates of the pen, so the goal is to train the weights so that z(t) closely matches the target, f (t). The learning algorithms are first allowed to learn for ten repetitions of the task. As a diagnostic, the error signals are not computed and the weights are frozen for a further five repetitions. This provides a way to check the accuracy of the final solution, demonstrating whether or not the algorithm has converged to an accurate solution. During this testing phase, feedback to the system comes from the true solution (Sussillo & Abbott, 2009). Specifically, Qz is replaced by Qf in Eq. (1). This avoids a drift in the phase of the solution that otherwise occurs when weights are frozen (see below). Additionally, for SUPERTREX, the exploratory pathway was shut off during these last five repetitions (z1 set to zero) to test how well the mastery pathway converged. This simple task is well suited to FORCE, which requires a known target, f , in order to compute the fully supervised error signal, e = z − f . FORCE was able to quickly find the correct solution to the task and maintained the correct result even after weights were frozen (Fig. 2B,C). Another measure of convergence is the activity of weight matrix, W , which quickly converged then stabilized (Fig. 2Cii, bottom). Error also remained low after learning was disabled (Fig. 2B). In summary, as expected, FORCE learned this task quickly and accurately. To apply RMHL and SUPERTREX to this task, we set e = (cid:107)z − f(cid:107)2 where (cid:107) · (cid:107) is the Euclidean norm, i.e. the distance of the pen from its target. This error contains strictly less information than the fully supervised error used by FORCE. RMHL performed well during learning, but the performance after weights were frozen (Fig. 2B,D) along with the cyclical changes in (cid:107)W(cid:107) during learning (Fig. 2Dii) demonstrate that the RMHL algorithm never actually converged. Instead, RMHL relied on rapid changes in W to mimic the correct output at each time point without truly learning it. Even when the number of learning trials was dramatically increased (to 100, not shown), RMHL's W continually oscillated rather than converging. SUPERTREX performed well on this task. During learning SUPERTREX performed slightly worse than FORCE and similar to RMHL (Fig. 2B,E). Unlike RMHL, though, SUPERTREX 7 Figure 2: Performance of three learning algorithms on Task 1. A) Task 1 is to draw a butterfly curve by directly controlling the x- and y-coordinates of a pen. Specifically, the outputs of the reservoir is z(t) = (x(t), y(t)) where x(t) and y(t) are the Cartesian coordinates of the pen. B) Euclidean distance of pen from target for FORCE (orange), RMHL (green), and SUPERTREX (purple). Learning was halted by freezing weights and exploration after ten periods, so the remain- ing five periods represent a testing phase. Ci) Target butterfly curve (red) versus the butterfly drawn by FORCE (blue) during the testing phase. Cii) Target (red) and actual (blue) outputs, x(t) and y(t), and the norm of the weight matrix, (cid:107)(W W T ) 2(cid:107)2, produced by FORCE. Di-ii) Same as C, but for RMHL. Di-ii) Same as C, but for SUPERTREX and the norm of each matrix, W1 (exploratory; green) and W2 (mastery; purple) are plotted separately. Vertical gray bar indicates the time at which weights were frozen. Note exploratory weight do change, but primarily at the start and is hard to see over the full trial timescale. See figure 3 for more details. 1 8 CiATask 1LearningTestingCii1sFORCEDiDiiRMHLEiEiiSUPERTREXDistance from Targetx(t)y(t)Wx=z1y=z2OutputTarget0.0010.010.1x(t)y(t)Wx(t)y(t)WRMHLFORCESUPERTREXExploratoryMasteryB Figure 3: The dynamics of SUPERTREX with an abruptly changed target. A) Target (red) and actual (blue) output. Same as Task 1, but the target was changed from a butterfly to a circle after ten periods. B) Detail around the time of change. Same colors as A (red for target and blue for total output), with addition of exploratory (green) and mastery (purple) components of the total output. Note that exploratory + mastery = total output. continued to track the target after learning was disabled, and performed similarly to FORCE during that phase (Fig. 2B). This, combined with the apparent convergence of (cid:107)W(cid:107) during learning (Fig. 2Eii, purple and green curves converge), indicates that the SUPERTREX algorithm did converge - albeit more slowly than FORCE. Interestingly, SUPERTREX produced less error during the testing phase than during learning (Fig. 2B). This is because exploration introduces random errors during learning, but exploration was turned off during testing so that output was produced only by the well-trained mastery pathway. This is comparable to findings in song birds in which natural or artificial suppression of neural activity in brain areas homologous to the basal ganglia reduce exploratory song variability and vocal errors (Kao et al., 2005). Looking at Fig. 2, it can be hard to tell whether the exploratory pathway is active, since the weights do not seem to change. This is due to the large timescale of the trial compared to the exploratory dominated phase, which only occurs as the algorithm is first adjusting to the task. An interesting illustration of the exploration / mastery handoff in SUPERTREX is provided by suddenly changing the target from a butterfly to a circle during learning (Fig. 3). The relative contributions from the exploratory pathway and the mastery pathway show that the exploratory pathway initially tracks the new target (Fig. 3B). Since the exploratory pathway is equivalent to RMHL, we know from above that the pathway is only mimicking the output through rapid weight changes. Over time, the mastery pathway learns from the activity of the exploratory pathway and begins taking over the generation of the output. This "handoff" from the exploratory to the mastery pathway produces a damped oscillation around the target (Fig. 3B). Task 2: Generating an unknown target from a scalar error signal Task 1 is a simple introductory task to compare the three learning algorithms, but it is also unrealistic in some ways that play towards FORCE's strengths. Specifically, the task involves producing an output, z, to match a known target, f , and error is computed in terms of the difference between z and f . In many tasks, motor output has indirect effects on the environment and the target and error are given in terms of these indirect effects. For example, consider a human or robot performing a drawing task. Motor output does not control the position of the pen directly, 9 x(t)y(t)A1s1mstotal outputtargetchangetargetexploratorymasteryBx(t)y(t)changetarget Figure 4: Performance of RMHL and SUPERTREX on Task 2. A) Task 2 is to draw the same butterfly (red) as in Task 1, but the reservoir output now controls the arm joint angles, z1(t) = θ1(t) and z2(t) = θ2(t). Error is still computed in terms of pen coordinates. FORCE is not applicable to this task. B) Euclidean distance of pen from target for RMHL (green) and SUPERTREX (purple). Learning was halted by freezing weights after ten periods, so the remaining five periods represent a testing phase. C-D) Same as Fig. 2D-E except that angles, θ1(t) and θ2(t) are plotted in place of pen coordinates. but instead controls the angles of the arm joints, which are nonlinearly related to pen position. On the other hand, error might be evaluated in terms of the distance of the pen from its target. Task 2 models this scenario. The goal in Task 2 is to draw the same butterfly from Task 1, parameterized by the same target coordinates f (t) = (x(t), y(t)). However, the reservoir output controls the angles of two arm joints (Fig. 4A), z(t) = (θ1(t), θ2(t)). We assume that the subject does not have access to the target angles that draw the butterfly. Instead, they only have access to the target pen coordinates, f , and its distance to the actual pen coordinates, which are related to the angles through a nonlinear function, h(θ1, θ2) = (x, y). FORCE cannot be applied directly to this task since the fully supervised error required for FORCE would need to be computed in terms of target angles instead of target pen position. RMHL and SUPERTREX can be applied to this task since they only require a signal that pro- vides enough information to determine whether error recently increased or decreased. In particular, 10 ATask 1LearningTesting1sCiCiiRMHLDiDiiSUPERTREXDistance from TargetOutputTarget10.010.1WRMHLSUPERTREXExploratoryMasteryBz =θ11z =θ22θ1(t)θ2(t)Wθ1(t)θ2(t) this is accomplished by setting e = (cid:107)h(z) − f(cid:107)2 where h(z) = (x, y) is the pen position. Once again, the task is divided into 10 learning cycles and 5 test cycles, with learning algorithms and the exploratory pathway of SUPERTREX disabled during the test cycles. Since the target angles, z(t), are unknown, feedback during testing cannot be replaced by the target as was done for Task 1. Instead, it is provided by the output from five previous periods. RMHL performed poorly on this task. It eventually mimicked the target (Fig. 4B), but once again failed to converge (Fig. 4B,C). SUPERTREX was able to track the target, and continue to produce it even after weight changes ceased (Fig. 4B,D). Hence, the combination of FORCE-like learning and RMHL-like learning implemented by SUPERTREX is able to learn a task that neither FORCE nor RMHL can learn on their own. Task 3: Learning and optimizing a task with multiple candidate solutions While FORCE cannot be applied to Task 2 as it is currently defined, it could be applied if the inverse of h were explicitly computed off-line to provide the target angles, (θ1, θ2) = h−1(f ), from which to compute a fully supervised error signal. This approach assumes that the subject knows the inverse of h and therefore does not easily extend to learning tasks in which h is difficult or impossible to invert. We now consider a task in which the error is not an invertible function of the motor output. Specifically, we consider an arm with three joints (Fig. 5A) and a cost function, C(θ(cid:48) 2, θ(cid:48) 3), that penalizes the movement of some joints more than others. Here, θ(cid:48) j is the time-derivative of θj. SUPERTREX can work with any penalty structure, making the choice arbitrary. Given that, we decided to loosely model our arm on a real human arm, with the joints corresponding to shoulder, elbow, and wrist. The penalties are larger for the angles controlling larger arm lengths, so the cost is lowest for the wrist joint, θ3, and largest for the shoulder joint, θ1, based on the fact that you are more likely to move your wrist than your entire shoulder and arm for a small reaching task - this can also be seen as an energy conservation principle, with larger costs associated to the more costly shoulder joint compared to the wrist joint. 1, θ(cid:48) For this task, there are infinitely many candidate solutions that successfully draw the butterfly, differing by the cost of joint movement. This turns our learning task into an optimization problem. Using FORCE for this task does not make sense, as it would require prior, explicit knowledge of the desired time series of joint angles. Essentially, it would require that the optimization problem had already been solved offline. RMHL and SUPERTREX can be applied to this problem by setting e = (cid:107)h(z) − f(cid:107)2 + C(θ(cid:48) 1, θ(cid:48) 2, θ(cid:48) 3). In this context, RMHL and SUPERTREX work as greedy search algorithms that make local changes to the angular output to reduce error and cost. Note, however, that the solution they find may not be globally optimal. We applied RMHL and SUPERTREX to this task using the same protocol for the learning and testing phases that we used for Task 2. RMHL performed poorly on this task (Fig. 5B,C), which is not surprising considering its poor performance on Task 2. SUPERTREX performed much better 11 Figure 5: Performance of RMHL and SUPERTREX on Task 3. A) Task 3 is to draw the same butterfly (red) as in Tasks 1 and 2, but the reservoir output now controls three arm joint angles, z1(t) = θ1(t), z2(t) = θ2(t), and z3(t) = θ3(t) with a different cost function associated to moving each joint. Error is computed in terms of pen coordinates and cost of moving joints. FORCE is not applicable to this task. B) Euclidean distance of pen from target (top) and cost (bottom; C(θ(cid:48) 3)) for RMHL (green) and SUPERTREX (purple). C,D) Same as Di and Ei in Fig. 2. E) Angular outputs and distance from target across two different SUPERTREX trials. The overall solution found was similar, with a mirrored rotation in one joint angle. 2, θ(cid:48) 1, θ(cid:48) 12 ATask 3LearningTesting1sCRMHLDSUPERTREXOutputTarget0.01RMHLSUPERTREXBz =θ11z =θ22z =θ3310-50.110.00010.0010.01DistanceCostθ10.11Distance1s0.01θ2θ3ETrial 1Trial 2 than RMHL: It was able to track the target, and continue to produce it even after weights were frozen (Fig. 5B,D). Over multiple runs SUPERTREX will find different solutions, as seen in Fig 5 E). The solution found will primarily depend on the initial condition, but the randomness in searching will also play a role. In this task, SUPERTREX tended to find similar solutions, except for random mirroring of certain angles. In summary, SUPERTREX can solve motor learning tasks in which there are multiple "correct" solutions with different costs. Disrupted learning as a model of Parkinson's disease The design of SUPERTREX was motivated in part by observations about the role of the basal ganglia in motor learning and Parkinson's disease (PD). PD is caused by the death of dopamine producing neurons in the basal ganglia, resulting in motor impairment. A common treatment for PD is a lesion of basal ganglia output afferents. Such lesions alleviate PD symptoms and impair performance on new learning tasks more than well-learned tasks (Obeso et al., 2009; Turner & Desmurget, 2010). These and other findings have inspired a theory of motor learning in which the basal ganglia are responsible for early learning, but not the performance of well-learned tasks and associations (Turner & Desmurget, 2010; H´elie et al., 2012). SUPERTREX is consistent with this theory if the exploratory pathway is interpreted as a basal ganglia pathway and the mastery pathway the cortical pathway. To test this model, we next performed an experiment in SUPERTREX that mimics the effects of PD and its treatment with basal ganglia lesion. The hand-off of learning from the exploratory to mastery pathway occurs extremely quickly in SUPERTREX due to the powerful, but biologically unrealistic RLS learning rule used in the mastery pathway (see, e.g., Fig. 3). To make SUPERTREX more biologically plausible for this experiment, we replaced the RLS learning rule with a least-mean-squares (LMS) rule by replacing P in Eq. (2) with the identity matrix (Hoerzer et al., 2014). This modified rule is more realistic because it avoids the complicated computation of the matrix, P , it makes the learning rules local, and it causes the mastery pathway to learn more slowly, which slows the hand-off from the exploratory pathway. We applied this modified SUPERTREX algorithm to Task 1. For 100 trials, learning proceeded normally. SUPERTREX learned the target more slowly than in Fig. 2 and with a slight degradation in performance due to the use of LMS instead of RLS learning in the mastery pathway (Fig. 6, early and late learning). This phase models normal learning before the onset of PD. By the end of this phase (Fig. 6, late learning), the task has become "well-learned" in the sense that output is generated by the mastery pathway instead of the exploratory pathway. The system output depending primarily on the mastery rather than exploratory pathway can be seen in Fig 6 C). For the next five trials, we corrupted the error signal to model the effects of PD. Since e models the error or cost of motor output, it is negatively related to dopamine release. Specifically, e ∝ Dmax − D where D quantifies dopamine release and Dmax is the maximum possible value of D. Hence, PD-induced dopamine depletion is modeled by artificially increasing e, which we achieve by setting e = (cid:107)f − z(cid:107)2 e = e + p where p(t) increases over time. Here, p = 0 corresponds to a healthy subject and, as p increases, SUPERTREX falsely evaluates more of its actions as being in error or costly. For our Parkinsonian task, we chose a p(t) that linearly increased to .1 over the duration of the corrupted learning phase. 13 Figure 6: SUPERTREX with a corrupted error signal models Parkinson's disease and its treatment. A) Euclidean distance of pen from its target for a modified version of SUPERTREX on Task 1. Learning proceeded normally for 100 trials. First five trials (early learning) and last five trials (late learning) are shown. The error signal was corrupted over the following five trials (corrupted learning) and the exploratory pathway was lesioned for the last five trials (post-lesion). B) Target (red) and actual (blue) outputs. C) Normed outputs from mastery pathway (purple) and exploratory pathway (green). D) Target butterfly (red) versus drawn curve during each of the plotted groups of five trials. 14 Distance from TargetABD0.010.10.6early learninglate learningcorrupted learningpost-lesiony(t)x(t)1sOutput Norm01C Even though the mastery pathway had taken over motor output before the error signal of the exploratory pathway was corrupted, the perceived increase in error caused the exploratory path- way to take over during the corrupted learning phase because the contribution of the exploratory pathway increases with error. Although the actual disruption may seem small (See Fig 6 C), where the exploratory activity is similar to that of early learning), the mismatch between actual error and perceived error during the corrupted learning phase results in highly inaccurate motor output (Fig. 6, corrupted learning phase) as activity leaves the learned manifold and is unable to recover. These results model the motor impairments associated with PD. Indeed, PD symptoms are believed to be caused, at least in part, by aberrant learning in the basal ganglia (Turner & Desmurget, 2010; Ashby et al., 2010). In the last five trials, we disabled the exploratory pathway, modeling basal ganglia lesion, and the feedback term, Qz, was replaced by Qf in Eq. (1) (see below and Discussion). SUPERTREX recovered nearly correct output during this last stage (Fig. 6, post-lesion phase) because the output had been stored in the mastery pathway before learning in the exploratory pathway was corrupted. As shown in Fig. 6, immediately before corruption began the mastery pathway was essentially solely responsible for generating the correct output. After the Parkinsonian effect, the final output is given solely by the mastery pathway as the malfunctioning exploratory pathway is lesioned. Thus, any degradation in the drawn butterfly is due to harmful changes made to the mastery pathway during the Parkinsonian effect. There are two main reasons why these harmful changes should be small. One is that the exploratory pathway changes are only kept if they result in a decrease in error even after taking into account the additional Parkinsonian error, or that the Parkinsonian error term makes changes due to exploration less likely to be accepted. Additionally, for sufficiently large errors the SUPERTREX component that controls transfer from exploration to mastery pathways shuts down, limiting the degree to which harmful perturbations can be assimilated. Thus, post- lesion performance will depend on the specific p(t) Parkinsonian effect used, along with the overall duration of the Parkinsonian effect. State Information Promotes Stability of Learned Output During our previous examples comparing FORCE, RMHL, and SUPERTREX, the comparison was made by allowing 10 trials of training the algorithm, and then with 5 trials of the learning algorithm shut off (weights frozen) to see if the method had converged. During this testing phase, feedback was modified. In task 1 it was replaced with the correct output (the target), and for tasks 2 and 3 it was replaced with the output from previous trials during learning, which nearly matched the target due to the learning algorithm being active. This allowed us to check whether an algorithm had converged, in the sense that there would be no further feedback and weight changes. However, providing the correct answer as feedback, also known as teacher forcing, could be considered cheating here. Teacher forcing essentially ignores stability of the solution and instead only checks whether the system can correctly produce the next time step of the solution given a perfect fit to the current time step. In order to address this, we repeated task 1, but without teacher forcing. FORCE has previously been shown to perform well in the absence of teacher forcing (Sussillo & Abbott, 2009; Abbott, Depasquale, & Memmesheimer, 2016), but it failed in our simulations (Fig. 7A, solid orange). We suspected that this was due to the extra additive noise, , added 15 Figure 7: Including target information in feedback promotes stability without teacher forcing. A) Euclidean distance of pen from its target for FORCE (orange), RHML (green), and SUPERTREX (purple) on Task 1. Same as Fig. 2B except feedback during the testing phase was not replaced with the true solution (teacher forcing), but is instead given by Qz exactly. B) Same as A, but for Task 2 and without FORCE (since it cannot be applied to Task 2). C) Butterfly drawn by SUPERTREX from the simulation in B. D-F) Same as A-C, except feedback was augmented by the target, Q[x y f ]. during learning. Noise is not typically included in applications of FORCE, but reservoir learning is known to be sensitive to noise and other perturbations (Vincent-Lamarre, Lajoie, & Thivierge, 2016; Sussillo, 2014; Miconi, 2017), which are ubiquitous in biological neuronal networks. Indeed, FORCE performed better when this noise was removed (Fig. 7A, dashed orange). Noise is an inherent part of RMHL and SUPERTREX, so they cannot be tested without it. Unsurprisingly, RMHL and SUPERTREX also perform poorly without teacher forcing (Fig. 7A,B,C). In summary, learning a noisy version of the target prevents all three algorithms from reproducing the target post-learning in the absence of teacher forcing. We resolve this issue by augmenting the feedback to include full information about the state of the system, allowing the system to self-correct. Specifically, we concatenated the x- and y- coordinates of the target pen position onto the feedback signal, replacing the Qz term in Eq. (1) with Q[z f ], during both training and testing. Under this modified framework, we again tested all three algorithms on Task 1 and tested RMHL and SUPERTREX on Task 2. For Task 1, this 16 Distance from Target10.010.1LearningTestingRMHLFORCESUPERTREX10.010.10.001LearningTestingTask 1Distance from Target10.1Task 20.0110.10.011sFeed Back Output OnlyFeed Back Output and TargetABDECF change is analogous to teacher forcing (since the target coordinates are the same as the target reservoir output). For Task 2, it is distinct from teacher forcing because the feedback is in terms of the Cartesian coordinates of the target, whereas the output must be in terms of arms' angles. Hence, for Task 2, the system must learn to self correct: If z and f differ then the networks need to learn how to generate the correct θ1 and θ2 to correct the error. This change greatly improved accuracy of FORCE and SUPERTREX, but not RMHL (Fig. 7D,E,F). Note that this change is not the same as just providing the correct answer as teacher forcing does. Teacher forcing essentially "resets" the system to be correct after every time step by replacing Qz with Qf , preventing drift. The augmented feedback instead provides sufficient information for the system to be autonomously self-correcting and the feedback is provided as-is with no context. In task 2, the algorithm does not have access to the solution it must produce (in terms of arm angles), but only has access to the target pen coordinates, which are non-linearly related to arm angles. This is akin to including a sensory feedback term, where the algorithms have sensory information about the actual and target positions, but do not have explicit information on necessary joint movements to make them overlap. Note that simply replacing the feedback from position to target will not result in convergence, e.g. replacing Qz with Qf throughout training and testing does not work. Both pieces of information together are required to build a stable system. The extra feedback term can be simplified further by changing f into a simple phase variable, which gives similar results as those shown in Fig. 7D,E,F (data not shown). Similar approaches have been proposed previously (Vincent-Lamarre et al., 2016). These approaches can model the presence of time-keeping neural populations. For example, in songbird, motor learning is believed to be supported by a timekeeping signal from HVC, which is extensively used in models of songbird learning (Doya & Sejnowski, 1995; Fiete, Fee, & Seung, 2007; Fee & Goldberg, 2011). Reward modulated learning with velocity control In all examples considered so far, the output of the reservoir controlled the position of a pen or the angle of arm joints. In control problems, motor output controls velocity or acceleration (e.g. applied force) of limbs or joints. From a naive perspective, SUPERTREX should still be able to complete such a task - random perturbations still change error, and SUPERTREX can learn to produce perturbations associated with lower error. However, a more careful consideration reveals that SUPERTREX and RMHL applied directly to control velocity would not be effective. To understand why, we first review and schematicize how SUPERTREX and RMHL successfully learn Task 1 where the output controls the position of the pen, then consider why they would not work when the reservoir output control the velocity of the pen. In Task 1, suppose the pen is displaced from its target (Fig. 8 top left) and an exploratory perturbation is made to the reservoir output that successfully moves the pen closer to its target (Fig. 8 bottom left). In this case, the change in error is negative (∆e ≈ e < 0), so the perturbation is correctly rewarded (see Eqs. (4) and (5)). Now consider Task 1 except that the reservoir output controls the velocity of the pen instead of the position. Again, suppose the pen is displaced from its target and also suppose that it is moving away from the target (Fig. 8 top middle). A beneficial exploratory perturbation changes the velocity of the pen in the direction of the target (Fig. 8 bottom middle). However, if the 17 Figure 8: Velocity Control in SUPERTREX. A) When the position of a pen is controlled and error is the distance of the pen from its target, a beneficial perturbation correctly results in a decreased error. B) When the velocity of the pen is controlled and error is computed in the same way, there are situations where a beneficial perturbation results in increased error. C) With velocity control, replacing the error by the derivative of the distance causes beneficial perturbations to correctly produce decreased error. perturbation was not strong enough to change the direction of the pen, then the error (which is measured as the distance of the pen from its target) will still have increased after the perturbation (as in Fig. 8 bottom middle), so that ∆e ≈ e > 0 and this perturbation will be penalized instead of rewarded (as again indicated by Eqs. (4) and (5)). This problem is overcome by taking the derivative of the error, specifically defining e to be the derivative of the distance between the pen and its target. When this change is made, then a reservoir controlling pen velocity will be correctly rewarded for beneficial perturbations Fig. 8 right) and penalized for harmful perturbations. To test these conclusions, we repeated Task 1 with SUPERTREX, except with output now corresponding to velocity rather than position, d[x, y] dt = z1 + z2 and we set [x(0) y(0)] = f(0). During the course of training this model, we discovered two other adjustment were required. As our goal was to track a signal, rather than reach a target, adding a penalty term based on velocity was helpful in order to prevent oscillations around our target, e.g. e = ∆((cid:107)f − z(cid:107)2 + γdtz) Unfortunately, we did not find a systematic way to determine γ. Instead, γ is chosen via iteration in order to prevent over- or under-damped oscillatory behavior. 18 Δe<0e = P-TΔe>0PerturbationΔe<0VelocityPositionTargetPerturbationPositionTargetPerturbationtime=ttime=t+dtPositional ControlRegular ErrorVelocity ControlRegular ErrorVelocity ControlDerivative Errore = P-Te = ΔP-TVelocityPositionTarget222ABC Figure 9: Velocity Control in SUPERTREX. A) A schematic of the velocity control task, which is identical to Task 1, except the velocity of the pen is controlled by the reservoir instead of its position. B) Using the regular error (distance of pen from target) produces large errors, but using the derivative of the distance produces smaller errors, especially during testing. C) Butterfly drawn during the testing phase using regular error and D) derivative error. 19 CATask 1LearningTesting1sRegular ErrorDistance from TargetxyOutputTarget0Regular ErrorDerivative ErrorB0.10.2z 1(t)=x'(t)z 2(t)=y'(t)DDerivative Error Additionally, standard feedback Qz clearly does not provide enough information - if we don't explicitly know our starting position, only knowing velocity does not help. Instead, we provided full-state information Q[x y f] since we care more about our position than our velocity in terms of feedback. This is also more realistic - it makes sense to modify z based on our position, rather than velocity, and position is more likely to be available as sensory feedback. Making these changes, we can compare SUPERTREX with error computed as the distance between the pen and its target (Fig. 9, "regular error") and with error computed as the derivative of the distance between the pen and its target (Fig. 9, "derivative error"). As predicted, SUPERTREX with velocity control performs better when using the derivative of the distance as the error signal (Fig. 9, compare red to purple in panel B, and compare panel C to D). Discussion We presented a novel, reward-modulated method of reservoir computing, SUPERTREX, that per- forms nearly as well as fully supervised methods. This is desirable as there are a broad class or problems where traditional supervised methods are not applicable, such as Tasks 2 and 3 that we considered. Moreover, humans can learn motor tasks from reinforcement signals alone (Izawa & Shadmehr, 2011). In place of a supervised error signal, SUPERTREX bootstraps from a dopamine- like, scalar error signal to a full error signal using rewarded exploration. This serves as an approx- imate target solution which is then transferred to a more traditional reservoir learning algorithm. This transfer of learned behavior to a mastery pathway, along with continued rewarded exploration, automatically creates a balanced system where the total output is correct, but the composition shifts over time from exploration to mastery. SUPERTREX performed similarly to FORCE on tasks where both were applicable, but also worked well on tasks where FORCE was not applicable. SUPERTREX also outperformed RMHL, a previously developed reward-modulated algorithm, on all tasks we considered. Unlike RMHL and other reinforcement learning models, SUPERTREX models the complemen- tary roles of cortical and basal ganglia pathways in motor learning. Under this interpretation, dopamine concentrations play the role of the reward signal, and the basal ganglia is the site of the RMHL-like, exploratory learning. Direct intra-cortical connections would then learn from Hebbian plasticity in the mastery pathway. Consistent with this interpretation, SUPERTREX produces inaccurate motor output when the reward signal is corrupted, modeling dopamine depletion in PD, but recovers the generation of well-learned output when the exploratory pathway is removed, model- ing basal ganglia lesions used to treat PD. Hence SUPERTREX provides a model for understanding the role of motor learning in PD and its treatments. As models of motor learning, reward-modulated algorithms like SUPERTREX and RMHL as- sume no knowledge of the relationship between motor output and error. In contrast, fully supervised algorithms like FORCE require perfect knowledge of this relationship. In reality, we learn through some combination of supervisory and reward-modulated error signals (Izawa & Shadmehr, 2011). To account for this, SUPERTREX could potentially be extended to incorporate both one-dimensional reward and higher-dimensional sensory feedback. The FORCE-like learning algorithm used for the mastery pathway of SUPERTREX is biolo- gistically unrealstic in some ways. The presence of the matrix, P , causes the rule to be non-local. However, we showed that SUPERTREX still works when P is removed to implement a local, LMS 20 learning rule (Fig. 6). Indeed, one can replace the mastery pathway with any supervised learning rule. This could open the way for an implementation of SUPERTREX with spiking neural net- works using existing supervised learning rules (Maass et al., 2002; Bourdoukan & Deneve, 2015; Abbott et al., 2016; Pyle & Rosenbaum, 2017). In order to have a fully spiking-based version of SUPERTREX, this would also require a spike-based reinforcement learning rule, most likely an eligibility-trace based rule (Seung, 2003; Xie & Seung, 2004; Fiete & Seung, 2006; Miconi, 2017). As with most other reservoir computing algorithms, SUPERTREX implements online learning in which a local error signal is provided and used at every time step. This is partly by design - SUPERTREX learns extremely (even unrealistically) quickly as weights are updated at a high frequency. This learning is slowed by some extent by switching to the more realistic LMS learning rule (as in Fig. 6). For some biological learning tasks, however, error signals are temporally sparse or reflect temporally non-local information. Trial-based learning rules for reservoir computing (Fiete & Seung, 2006; Miconi, 2017) are applicable in the presence of sparse or non-local rewards. At least one of these algorithms learns very slowly, requiring thousands of trials (Miconi, 2017), which may be an inevitable consequence of learning from sparse rewards. In reality, biological motor learning likely makes use of both online and sparse feedback. An extension of SUPERTREX that accounts for both types of feedback could be more versatile and realistic. SUPERTREX is conceptually an extension of SPEED (Ashby et al., 2007), which has a similar framework for categorization and other discrete tasks. SPEED learns to map arbitrary discrete inputs to discrete outputs, such as in categorization tasks. While the architecture and learning rule are similar to SUPERTREX, SPEED cannot produce continuous, dynamical output and requires a separate pathway for each possible input-output pairing. SUPERTREX could also be compared to a class of RNN algorithms that use a teacher network to train the final output network. However, many of these networks use the activity of the teacher network as a way to train the recurrence J of the output network; in SUPERTREX, there is only one recurrent network (used for both outputs). These methods are often even more biologically implausible - for example, the recent FULL-FORCE extension of FORCE (DePasquale, Cueva, Rajan, Abbott, et al., 2018) feeds the target signal info the first, chaotic reservoir, and then uses the activities of each reservoir unit in the teacher network as a target for training the second network, drastically increasing the amount of supervision required. SUPERTREX loses accuracy when learning is halted when feedback consists solely of the sys- tem's output (Fig. 7A-C) due to the fact that it learns from a noisy estimate of the target. This shortcoming can be overcome by augmenting the feedback with the target, allowing the system to learn to self-correct noise-induced errors (Fig. 7D-F). FORCE is susceptible to the same instabili- ties as SUPERTREX under the biologically realistic assumption of noise during learning (Fig. 7A), but SUPERTREX can solve tasks that FORCE cannot (Figs. 4 and 5). RMHL is also susceptible to the same instabilities and is applicable to the same tasks as SUPERTREX, but the instabil- ities in RMHL are not resolved by including target information in the feedback as they are for SUPERTREX (Fig. 7C,D). Hence, SUPERTREX is the only one of the three algorithms that can be applied to reward-modulated learning tasks and achieves stability with target information in the feedback. Stability in reward-modulated reservoir computing without target information in the feedback term remains an open problem. This problem could potentially be solved by providing external input in-phase with the target output. This could help the reservoir "keep time" by re- aligning the reservoirs' state on each trial, allowing the system to self-correct its phase. A similar 21 approach was shown to improve robustness of FORCE to perturbations in previous work (Vincent- Lamarre et al., 2016). Interestingly, biology may have already solved this problem. Research by Toledo-Suarez, Duarte, and Morrison (Toledo-Su´arez, Duarte, & Morrison, 2014) has found that the striatum may act as a reservoir computer that processes state information. Rather than rely on raw inputs, the motor learning system instead has access to pre-processed state information that is both simpler and more relevant. In SUPERTREX, this could correspond to replacing our simple feedback of raw state information Qz or Q[zf ] with Qs, where s is a pre-processed state information vector. s could even come from another reservoir, designed to ensure s contains maximally relevant information to the task at hand. This would be an interesting extension to SUPERTREX. In summary, SUPERTREX is a new biologically inspired framework for reservoir computing that is more realistic and more effective than its predecessors. Using a general error signal allows for SUPERTREX to be used in places where a more powerful algorithm like FORCE cannot. The hand off from exploration to mastery allows SUPERTREX to perform nearly as well as FORCE with the generality of reward-modulated algorithms. Moreover, SUPERTREX offers a computational formalization of widely supported theories of motor learning and reproduces several experimental and clinical findings. Hence, this new framework opens the way for a truly two-way communication between biological and computational theories of motor learning. Materials and methods Simulation and Reservoir Parameters All simulations were performed using a forward Euler method, with dt = 0.2ms. Each task period or "trial" was 104 ms long and all simulations except those in Fig. 6 had 15 trials. Fig. 6 had 110 trials. The reservoir equation used in all algorithms was = −x + Jr + Qz τ dx dt where r = tanh(x) + αη, η was uniformly drawn from [−1, 1] on every time step, τ =10, and α = 2.5 × 10−2 during training and α = 0 during testing. Reservoir size was set to N = 1000 neurons, with connection probability p = 0.1. Connection strengths in J were normally distributed with mean 0 and variance λ2/(pN ) with λ = 1.5. Feedback Q was dense, with weights uniformly between −1 and 1. Initial readout weights for RMHL and SUPERTREX exploratory pathway, as well as weights for FORCE and the SUPERTREX mastery pathway were initialized at 0. Initial voltages were set uniformly between −0.5 and 0.5, while initial rates were the hyperbolic tangent of initial voltages. Displayed outputs and errors were low pass filtered according to τM SE dM SE(t) dt dz(t) dt τbar = −M SE(t) + M SE(t) = −z(t) + z(t) where τM SE = 1000, τbar = 10, and x represents a low pass filtered version of the variable x. The 22 plotted "distance from target" was computed as the pen from its target. √ M SE where M SE(t) is the squared distance of FORCE Reservoir output was z = W r and the learning rule is τw dW dt = −[z − f ]rT P with τw = 0.02. The matrix P is a running estimate of the inverse of the correlation matrix of rates r, initialized to and updated according to P (0) = 1 γ I where τP = dt, γ = 10 is a constant and I is the identity matrix. The matrix P is only updated every 10 time steps in order to save on computing time. τp dP dt = − P r rT P 1 + rT P r RMHL For RMHL, outputs were given by and the learning rule was z = W r + Ψ(e)η τw dW dt = Φ(e)(z)rT where τw = 0.02, η is uniformly distributed noise between [−1, 1], and the high-pass filtered version, x, of variable x was computed as τ with τ = 1 used for all tasks and trials. SUPERTREX = −x + x dx dt x = x − x. Updates to P were identical to the method used in FORCE above. Relevant other changes are z1 = W1r + Ψ(e)η z2 = W2r z = z1 + z2 for η uniformly drawn from [−1, 1]. For the learning algorithm, τw τw dW1 dt dW2 dt = Φ(e)zrT = −kz1rTP 23 with τw remaining 0.02 and constant learning rate k which varies per task. Finally, an extra condition was imposed on updates to P , W2. Both updates were multiplied by (−0.5 ∗ tanh(5 × 105 ∗ (e − (1.5 × 10−3))) + 0.5), which acts as a soft threshold around e = 1.5 × 10−3. Effectively, for errors larger than this the mastery pathway would not activate. Performance was similar, but slightly slower, without this thresholding. Tasks In all tasks, the target was to draw a butterfly, given by a polar curve x(t) = r(t) cos(t) and y(t) = r(t) sin(t) where r(t) =c[9 − sin(qt) + 2 sin(3qt)+ 2 sin(5qt) − sin(7qt) + 3 cos(2qt) − 2 cos(4qt)] and c = 1/ maxt[r(t)] is a normalizing constant. For a single repetition, t went from 0 to 104ms, and q = 2π 104 scales the system such that qt goes from 0 to 2π over the duration. In task 2, the task is instead to draw a butterfly by controlling two angles, representing radians from y axis, and radians from the first joint. The arm is positioned at (0,-2), and each arm segment has fixed length of 1.8. h(z) is therefore (cid:20) h(z) = 1.8 sin(z1π) + 1.8 sin((z1 + z2)π) −2 + 1.8 cos(z1π) + 1.8 cos((z1 + z2)π) (cid:21) (cid:20) In task 3, now there are three angles to control. The arm is positioned at (0,-2), and each arm segment has fixed length. The first segment has length 1.8, the next 1.2, and the final .6. h(z) is therefore h(z) = 1.8 sin(z1π) + 1.2 sin((z1 + z2)π) + .6 sin((z1 + z2 + z3)π) −2 + 1.8 cos(z1π) + 1.2 cos((z1 + z2)π) + .6 cos((z1 + z2 + z3)π) √ In the first task, Ψ(x) = 0.025 × 4 √ 10x and Φ(x) = −5 4 √ When testing for swapping targets, Ψ(x) = 0.1×(−5x)0.3 and Φ(x) = 2.5× 4 learning rate k was 0.5. √ For the second task, SUPERTREX learning rate k was still 0.5, Ψ(x) = 0.01 × 5 x for both RMHL and SUPERTREX. x. For SUPERTREX, 10x, and (cid:21) √ Φ(x) = 5 × 4 x. √ For the third task, SUPERTREX learning rate was k = 0.9, Ψ(x) = .025 × 4 √ 5 × 4 x. The error metric was changed slightly, to 10x, and Φ(x) = e = h(z) − f2 + α z1 + β z2 + γ z3 for α = 0.1, β = 0.05, γ = 0. This implemented an additional cost for moving joints; highest for the longest arm segment, and 0 for the smallest arm segment. For the corrupted learning example, LMS learning was used, which is obtained by setting P = I. The learning rate was changed to k = 0.003. Note that LMS learning rather than RMS learning generally requires a much lower learning rate. Other parameter values were the same as in the first task. The perturbation, p(t), increased linearly from 0 to 0.1 over the corrupted learning timeframe. 24 For the velocity controlled example in Fig. 8, more significant changes were needed. As detailed, and [x, y](0) = f (0) d[x, y] dt = z1 + z2 as well as using full state feedback Q[x y f]. Learning rate k was .025, as smaller velocities were needed relative to direct control of output. Velocity penalty γ = .3. Ψ and Φ were the same as in task 1. Finally, we changed how we we calculated e. Rather than use a high pass filter as a crude derivative estimator, we instead used a finite difference approximation e = e(t) − e(t − dt). Note that, as described above, e(t) now refers to ∆d(t) = d(t)− d(t− dt) where d(t) = (cid:107)f − z(cid:107)2 + γdtz, e.g. the squared euclidean distance between the position and target plus a penalty term. Thus, our total update metric e = d(t) − 2d(t − dt) + d(t − 2dt), or the finite difference approximation to the second derivative of our error metric, which is euclidean distance plus penalty. Acknowledgments This work was supported by National Science Foundation grants DMS-1517828, DMS-1654268, and DBI-1707400. The authors thank Jonathan Rubin, Robert Turner, and Robert Mendez for helpful comments. References Abbott, L. F., Depasquale, B., & Memmesheimer, R.-m. (2016). Building Functional Networks of Spiking Model Neurons. Nat. Neurosci., 19 (3), 1 -- 16. Andalman, A. S., & Fee, M. S. (2009). A basal ganglia-forebrain circuit in the songbird biases motor output to avoid vocal errors. Proc. Natl. Acad. Sci. U. S. A., 106 (30), 12518 -- 23. Aronov, D., Andalman, A., & Fee, M. (2008). A specialized forebrain circuit for vocal babbling in the juvenile songbird. Science (80-. )., 320 , 630 -- 635. Ashby, F. G., Ennis, J. M., & Spiering, B. J. (2007). A neurobiological theory of automaticity in perceptual categorization. Psychol. Rev., 114 (3), 632 -- 56. Ashby, F. G., Turner, B. O., & Horvitz, J. C. (2010). Cortical and basal ganglia contributions to habit learning and automaticity. Trends Cog. Sci., 14 (5), 208 -- 215. Bottjer, S. W., Miesner, E. a., & Arnold, a. P. (1984). Forebrain lesions disrupt development but not maintenance of song in passerine birds. Science (80-. )., 224 (4651), 901 -- 903. Bourdoukan, R., & Deneve, S. (2015). Enforcing balance allows local supervised learning in spiking recurrent networks. In Adv. neur. in. (pp. 982 -- 990). Brainard, & Doupe, A. (2000). Interruption of a basal gangliaforebrain circuit prevents plasticity of learned vocalizations. Nature, 404 . Brainard, & Doupe, A. (2002). What songbirds teach us about learning. Nature, 417 , 351 -- 358. Brainard, M. S. (2004). Contributions of the anterior forebrain pathway to vocal plasticity. Ann. NY Acad. Sci., 1016 (1), 377 -- 394. 25 Carelli, R. M., Wolske, M., & West, M. O. (1997). Loss of lever press-related firing of rat striatal forelimb neurons after repeated sessions in a lever pressing task. J. Neurosci., 17 (5), 1804 -- 1814. Churchland, M. M., Cunningham, J. P., Kaufman, M. T., Foster, J. D., Nuyujukian, P., Ryu, S. I., & Shenoy, K. V. (2012). Neural population dynamics during reaching. Nature, 487 (7405), 51 -- 6. DePasquale, B., Cueva, C. J., Rajan, K., Abbott, L., et al. (2018). full-force: A target-based method for training recurrent networks. PloS one, 13 (2), e0191527. Doya, K., & Sejnowski, T. J. (1995). A novel reinforcement model of birdsong vocalization learning. In Adv. neur. in. (pp. 101 -- 108). Fee, M. S. (2014). The role of efference copy in striatal learning. Curr. Opin. Neurobiol., 25 , 194 -- 200. Fee, M. S., & Goldberg, J. H. (2011). A hypothesis for basal ganglia-dependent reinforcement learning in the songbird. Neuroscience, 198 , 152 -- 170. Fiete, I. R., Fee, M. S., & Seung, H. S. (2007). Model of birdsong learning based on gradient estimation by dynamic perturbation of neural conductances. J. Neuropysiol., 98 (4), 2038 -- 2057. Fiete, I. R., & Seung, H. S. (2006). Gradient learning in spiking neural networks by dynamic perturbation of conductances. Phys. Rev. Lett., 97 (4), 048104. H´elie, S., Paul, E. J., & Ashby, F. G. (2012). A neurocomputational account of cognitive deficits in Parkinson's disease. Neuropsychologia, 50 (9), 2290 -- 302. Hennequin, G., Vogels, T. P., & Gerstner, W. (2014). Optimal control of transient dynamics in balanced networks supports generation of complex movements. Neuron, 82 (6), 1394 -- 406. Hoerzer, G. M., Legenstein, R., & Maass, W. (2014). Emergence of complex computational structures from chaotic neural networks through reward-modulated hebbian learning. Cereb. Cort., 24 (3), 677 -- 690. Izawa, J., & Shadmehr, R. (2011). Learning from sensory and reward prediction errors during motor adaptation. PLoS Comput. Biol., 7 (3), 1 -- 11. Jaeger, H., & Haas, H. (2004). Harnessing nonlinearity: predicting chaotic systems and saving energy in wireless communication. Science, 304 (5667), 78 -- 80. Kao, M. H., Doupe, A. J., & Brainard, M. S. (2005). Contributions of an avian basal ganglia -- forebrain circuit to real-time modulation of song. Nature, 433 (7026), 638 -- 643. Kawai, R., Markman, T., Poddar, R., Ko, R., Fantana, A. L., Dhawale, A. K., . . . Olveczky, B. P. (2015). Motor cortex is required for learning but not for executing a motor skill. Neuron, 86 (3), 800 -- 812. Laje, R., & Buonomano, D. V. (2013). Robust timing and motor patterns by taming chaos in recurrent neural networks. Nat. Neurosci., 16 (7), 925 -- 33. Lukosevicius, M., Jaeger, H., & Schrauwen, B. (2012). Reservoir Computing Trends. KI - Kunstliche Intelligenz , 26 (4), 365 -- 371. Maass, W., Natschlager, T., & Markram, H. (2002). Real-time computing without stable states: a new framework for neural computation based on perturbations. Neural Comput., 14 (11), 2531 -- 60. Mante, V., Sussillo, D., Shenoy, K. V., & Newsome, W. T. (2013). Context-dependent computation by recurrent dynamics in prefrontal cortex. Nature, 503 (7474), 78 -- 84. 26 Miconi, T. (2017). Biologically plausible learning in recurrent neural networks reproduces neural dynamics observed during cognitive tasks. E-Life, 6 . Miyachi, S., Hikosaka, O., & Lu, X. (2002). Differential activation of monkey striatal neurons in the early and late stages of procedural learning. Exp Brain Res, 146 (1), 122 -- 126. Miyachi, S., Hikosaka, O., Miyashita, K., K´ar´adi, Z., & Rand, M. K. (1997). Differential roles of monkey striatum in learning of sequential hand movement. Exp. brain res., 115 (1), 1 -- 5. Obeso, J. a., Jahanshahi, M., Alvarez, L., Macias, R., Pedroso, I., Wilkinson, L., . . . Rothwell, J. C. (2009). What can man do without basal ganglia motor output? The effect of com- bined unilateral subthalamotomy and pallidotomy in a patient with Parkinson's disease. Exp. Neurol., 220 (2), 283 -- 92. Olveczky, B. P., Andalman, A. S., & Fee, M. S. (2005). Vocal experimentation in the juvenile songbird requires a basal ganglia circuit. PLoS Biol., 3 (5), e153. Olveczky, B. P., Otchy, T. M., Goldberg, J. H., Aronov, D., & Fee, M. S. (2011). Changes in the neural control of a complex motor sequence during learning. J. Neurophysiol., 106 (1), 386 -- 97. Pasupathy, A., & Miller, E. K. (2005). Different time courses of learning-related activity in the prefrontal cortex and striatum. Nature, 433 (7028), 873 -- 876. Poldrack, R. A., Sabb, F. W., Foerde, K., Tom, S. M., Asarnow, R. F., Bookheimer, S. Y., & Knowlton, B. J. (2005). The neural correlates of motor skill automaticity. J. Neurosci., 25 (22), 5356 -- 5364. Pyle, R., & Rosenbaum, R. (2017). Spatiotemporal dynamics and reliable computations in recurrent spiking neural networks. Phys. Rev. Lett., 118 (1), 018103. Russo, A. A., Bittner, S. R., Perkins, S. M., Seely, J. S., London, B. M., Lara, A. H., . . . others (2018). Motor cortex embeds muscle-like commands in an untangled population response. Neuron, 97 (4), 953 -- 966. Seung, H. S. (2003). Learning in spiking neural networks by reinforcement of stochastic synaptic transmission. Neuron, 40 (6), 1063 -- 1073. Shenoy, K. V., Sahani, M., & Churchland, M. M. (2013). Cortical control of arm movements: a dynamical systems perspective. Annu. Rev. Neurosci., 36 , 337 -- 59. Sompolinsky, H., Crisanti, a., & Sommers, H. J. (1988). Chaos in random neural networks. Phys. Rev. Lett., 61 (3), 259 -- 262. Sussillo, D. (2014). Neural circuits as computational dynamical systems. Curr. Opin. Neurobiol., 25 , 156 -- 163. Sussillo, D., & Abbott, L. F. (2009). Generating coherent patterns of activity from chaotic neural networks. Neuron, 63 (4), 544 -- 557. Sussillo, D., Churchland, M. M., Kaufman, M. T., & Shenoy, K. V. (2013). A neural network that finds naturalistic solutions for the production of muscle activity. Nat. Neurosci., 18 (7). Tang, C. C., Root, D. H., Duke, D. C., Zhu, Y., Teixeria, K., Ma, S., . . . West, M. O. (2009). Decreased Firing of Striatal Neurons Related to Licking during Acquisition and Overtraining of a Licking Task. J. Neurosci., 29 (44), 13952 -- 13961. Toledo-Su´arez, C., Duarte, R., & Morrison, A. (2014). Liquid computing on and off the edge of chaos with a striatal microcircuit. Frontiers in computational neuroscience, 8 , 130. Turner, R. S., & Desmurget, M. (2010). Basal ganglia contributions to motor control: a vigorous tutor. Curr. Opin. Neurobiol., 20 (6), 704 -- 16. 27 Vincent-Lamarre, P., Lajoie, G., & Thivierge, J.-P. (2016). Driving reservoir models with os- cillations: a solution to the extreme structural sensitivity of chaotic networks. J. Comput. Neurosci., 41 (3), 305 -- 322. Xie, X., & Seung, H. S. (2004). Learning in neural networks by reinforcement of irregular spiking. Physical Review E , 69 (4), 041909. 28
1809.01051
1
1809
2018-09-04T15:42:00
Scaling Spike Detection and Sorting for Next Generation Electrophysiology
[ "q-bio.NC" ]
Reliable spike detection and sorting, the process of assigning each detected spike to its originating neuron, is an essential step in the analysis of extracellular electrical recordings from neurons. The volume and complexity of the data from recently developed large scale, high density microelectrode arrays and probes, which allow recording from thousands of channels simultaneously, substantially complicate this task conceptually and computationally. This chapter provides a summary and discussion of recently developed methods to tackle these challenges, and discuss the important aspect of algorithm validation, and assessment of detection and sorting quality.
q-bio.NC
q-bio
Scaling Spike Detection and Sorting for Next Generation Electrophysiology Ma(cid:138)hias H. Henniga, Cole Hurwitza and Martino Sorbaroa,b September 5, 2018 a Institute for Adaptive Neural Computation, School of Informatics, University of Edinburgh, Informatics Forum, 10 Crichton Street, Edinburgh EH8 9AB, Scotland, United Kingdom b School of Computer Science and Communication, KTH Royal Institute of Technology, Lindstedtsvagen 5, 114 28 Stockholm, Sweden Abstract Reliable spike detection and sorting, the process of assigning each detected spike to its originating neuron, is an essential step in the analysis of extracellular electrical recordings from neurons. (cid:140)e volume and complexity of the data from recently developed large scale, high density microelectrode arrays and probes, which allow recording from thousands of channels simultane- ously, substantially complicate this task conceptually and computationally. (cid:140)is chapter provides a summary and discussion of recently developed methods to tackle these challenges, and discuss the important aspect of algorithm validation, and assessment of detection and sorting quality. 1 Introduction Extracellular electrical recording of neural activity is an essential tool in neuroscience. If an elec- trode is placed su(cid:129)ciently close to a spiking neuron, the extracellular potential recorded o(cid:137)en contains a clear, readily detectable signature of the action potential. As extracellular electrodes do not interfere with neural function, such recordings provide an unbiased and precise record of the functioning of intact neural circuits. Recent progress in CMOS technology (Complementary metal-oxide semiconductor technology for low power integrated circuits) has provided systems that allow recording from thousands of closely spaced channels simultaneously with ever increasing density and sampling rates. With this technology, it becomes possible to reliably monitor several thousand neurons simultaneously both in vitro and in vivo (Eversmann et al., 2003; Berdondini et al., 2005; Frey et al., 2010; Ballini et al., 2014; Muller et al., 2015; Yuan et al., 2016; Lopez et al., 2016; Jun et al., 2017b; Dimitriadis et al., 2018). (cid:140)is is a signi(cid:128)cant advancement as it enables, for the (cid:128)rst time, the systematic investigation of interactions between neurons in large circuits. Understanding these interactions will contribute to learning more about how neural circuitry is altered by cellular changes in diseases, injury, and during pharmacological interventions. To appreciate the advantages of recording the activity of many neurons, it is important to em- phasize that neural circuits are usually highly diverse and heterogeneous (Hrom´adka et al., 2008; Buzs´aki and Mizuseki, 2014; Panas et al., 2015). Not only do they consist of di(cid:130)erent neuron types, 1 but even within groups of neurons of the same type, the (cid:128)ring rates may di(cid:130)er by orders of magni- tude. (cid:140)is observation has been made consistently in vitro and in vivo, and it stands to reason that this has biological relevance. Conventional technologies, which allow simultaneous recording of a handful (rarely more than a hundred) of neurons, severely under-sample highly heterogeneous populations. If the recorded neurons are not representative of the whole population, both exper- imental accuracy and reproducibility between experiments will be negatively a(cid:130)ected. Moreover, dense recording systems increase the fraction of neurons isolated in a local population, to a level that was, so far, only accessible with calcium imaging. A further advantage of recording many neurons at once is that it can be an e(cid:130)ective way of probing neural excitability and connectivity, using functional interactions as a proxy measure for the e(cid:130)ects of synaptic interactions. In vitro assays are particularly suited for investigation of functional interactions, as they can be augmented with stimulation, (cid:131)uorescent labeling and targeted optogenetic stimulation (Zhang et al., 2009; Obien et al., 2015). A combination of dense multielectrode arrays and imaging technologies could allow phenotyping at the level of single cells, potentially in combination with further modalities such as gene expression pro(cid:128)ling. (cid:140)e high yield of such approaches thus provides entirely new possibilities for systematic assessment of the roles of di(cid:130)erent genotypes and of drug e(cid:130)ects. (cid:140)e analysis of single neuron activity requires the correct assignment of each detected spike to the originating neuron, a process called spike sorting. In this chapter, we will provide an overview of the most frequently employed methods for the spike sorting for large-scale, dense multielec- trode arrays. While many of the issues discussed will also apply to dense in vivo probes, the focus is on in vitro arrays, because they typically provide a large surface area evenly covered with recording channels, which is advantageous for spike sorting. A major additional challenge in in vivo recordings is tissue movement, which causes the signals of neurons to dri(cid:137) over time. For an excellent review of the challenges encountered in vivo, and of methodology for conventional recording devices with fewer channels, the reader may consult Rey et al. (2015). In the (cid:128)rst section, we will discuss in more detail the technical and practical issues that are introduced when moving from conventional devices with tens of channels to larger, more dense systems. Next, we will introduce the main components of modern spike sorting pipelines, and then discuss each component and existing algorithms in detail. Finally, we will provide an overview of approaches for validation of the quality of these algorithms. 2 Challenges for large-scale spike sorting On both conventional and high density recording devices, electrodes will usually pick up the ac- tivity of multiple neurons. While it is possible to directly analyse the multi-unit activity (MUA) from each channel, spike sorting is required to resolve single-unit activity (SUA). Spike sorting resembles the classic "cocktail party" problem: to isolate the voice of a single speaker in a crowd of people. Since the recorded spike waveforms di(cid:130)er in shape and amplitude among neurons, the resulting signal can be de-mixed using either dimensionality reduction paired with cluster- ing or spike templates along with template matching. (cid:140)ese approaches have been successfully employed on conventional devices with few, spatially well separated channels. On large-scale, dense arrays, however, these traditional methods become more di(cid:129)cult both computationally and algorithmically. Instead of (cid:128)nding a single voice in a crowd, the challenge is to isolate the voices of thousands of speakers in a room equipped with thousands of microphones. Overcoming this 2 challenge is imperative as wrong assignments can severely bias subsequent analysis of neuronal populations (Ventura and Gerkin, 2012). Spike sorting is a tractable problem for conventional extracellular recordings as it is commonly done for each recording channel separately. In this case, only a small number of neurons are ex- pected to contribute to the signal on each channel, which allows the use of precise, but computa- tionally more costly algorithms. Also, most existing algorithms for spike sorting still include an element of manual intervention to adjust or improve sorting results. (cid:140)ese traditional algorithms struggle when faced with large-scale, dense arrays. On dense arrays, a single action-potential from a neuron is visible on multiple, nearby chan- nels. As a result, spike sorting on single channels is no longer appropriate. Removing duplicate events is feasible in principle, but becomes challenging when nearby neurons are (cid:128)ring with high synchrony. Poor treatment of duplicate removal can lead to false exclusions of action potentials or retention of multiple spikes from the same action potential. Conventional spike sorting algorithms also struggle with the sheer volume of data large-scale arrays produce. For instance, a recording from 4,096 channels with 18kHz sampling rate yields about 140 megabytes per second, or over 8 gigabytes per minute. Simply reading this data volume from hard disk into memory for analysis can be a severe bo(cid:138)leneck in any spike sorting pipeline. In addition, the massive data volume prevents extensive manual curation of spike sorting results. Highly automated pipelines with minimal need for intervention are needed to overcome these challenges and to fully exploit the capabilities of dense arrays. 3 From raw data to single neuron activity A typical spike sorting pipeline begins with the detection of candidate events followed by some method of assigning these events to speci(cid:128)c neurons. (Lewicki, 1998; Rey et al., 2015). On large- scale arrays, two approaches have emerged as particularly suitable. One method is based on cre- ating spike templates and then performing template matching. (cid:140)e other method relies on feature extraction and clustering, using both the spike shape and estimated location of the event. A sum- mary of the steps required to obtain sorted spikes from raw data is shown in Figure 1. Each of these steps is discussed in more detail below. 3.1 Spike detection Spikes in the raw signal take the form of biphasic de(cid:131)ections from a baseline level. (cid:140)ey can be found through detection of threshold crossings and by using additional shape parameters such as the presence of a biphasic shape as acceptance criteria. As the noise levels may vary among channels and over time, the threshold is usually de(cid:128)ned relative to the noise level, which is esti- mated from portions of the raw signal that do not contain spikes. It is worth noting that signal (cid:131)uctuations in extracellular data are typically highly non-Gaussian. As a result, a noise estimate based on percentiles is more accurate and also easier to obtain, as opposed to computing the signal variance (Fee et al., 1996; Muthmann et al., 2015). (cid:140)e choice of the detection threshold determines which events are retained for further analysis. Spikes from well-detected neurons are easily identi(cid:128)able, but de(cid:131)ections with amplitudes closer to the background noise level are harder to isolate. Since there is typically no clear-cut separation between spikes and noise, events detected close to the threshold may originate from neurons for 3 Figure 1: Schematic overview of existing spike sorting pipelines for high density microelectrode arrays. Following detection, either neuron templates are formed based on the spatio-temporal event footprints and then used to detect these units in a second pass, or current sources are estimated and clustered together with waveform features, or a mask is created to restrict clustering to channels with a detectable signal. (cid:140)e output consists of a list of spike time stamps for each identi(cid:128)ed neuron, which o(cid:137)en has to be corrected in a (cid:128)nal manual curation step. which only an incomplete activity record can be obtained. (cid:140)e magnitude of electrical noise, which can be estimated when recording from an empty array, is usually much smaller than the magnitude of the (cid:131)uctuations recorded in the absence of clearly visible spiking activity (Muthmann et al., 2015). (cid:140)is indicates that a large component of the recorded signal (cid:131)uctuations are due to neural activity, such as neurons located further away from the electrode, or smaller events such as currents during synaptic transmission. (cid:140)e analysis of recordings from the retina shows that even very small detected signals may re(cid:131)ect activity that is typical of stimulus-evoked responses from retinal ganglion cells, hence carries signatures of neural activity rather than noise (Figure 2). As a result, the detection step signi(cid:128)cantly a(cid:130)ects the subsequent isolation of single neuron activity. Choosing a high detection threshold is not an ideal solution as this will potentially leave valid spikes undetected. In contrast, a low threshold guarantees reliable detection of neurons with stronger signals, but also increases the fraction of false positives. As a good compromise, a strategy can be adopted to detect events with a low threshold, and to subsequently discard unreliable units. (cid:140)is can either be done a(cid:137)er detection, for instance by using a classi(cid:128)er trained on true spikes and noise events obtained from channels not reporting neural activity (Hilgen et al., 2017), or by removing sorted units with a small number of spikes or poor clustering metric scores post spike sorting (Hill et al., 2011). Recently, a new method for spike detection using a pre-trained neural network was introduced by Lee et al. (2017). (cid:140)is method was shown to outperform conventional threshold-based methods on simulated ground truth data, in particular by achieving a lower false positive rate. When run on a modern GPU, a neural-network based method is also much faster. (cid:140)is is a very promising avenue, although the considerations regarding detection thresholds outlined above still remain 4 Feature extractionPC1PC2ConstructTemplatesby ClusteringEstimatespike locationsMethod #2Clusterlow dim.embeddingsMethod #1TemplatematchFlood filland maskSpike Detection Manualcuration Figure 2: Even small events detected on a high density array contain signatures of neural activity. A, Density plot of spatially binned spike counts, estimated from detected and spatially localised spikes using the method described by Hilgen et al. (2017). Spikes were recorded from a light-stimulated mouse retina. Spike detection was performed with a low threshold, hence false positives were registered in areas where no neural activity was recorded, such as the optic disk on the centre le(cid:137). B, No clear separation between spikes and noise is seen for recorded amplitudes, or for average amplitudes of units following spike sorting. C, Individual, randomly selected units with small (le(cid:137)) and large amplitudes (right) both show signatures of light stimulation during presentation of full (cid:128)eld (cid:131)ashes. For each unit, the light-evoked peri-stimulus time histogram (le(cid:137)), and examples and the average of spike waveforms (right) are shown. (cid:140)e recording was contributed by Gerrit Hilgen and Evelyne Sernagor, University of Newcastle. 5 02550Channels0204060123log10(Spike count/bin)0510Spike amplitude (z-score)0.00.51.01.5DensitySpikesUnits024Time (s)050Rate (Hz)100V1ms024Time (s)ABClight onlight offlight onlight offSmall spikesLarge spikes relevant. 3.2 Dealing with duplicate spikes Unlike in conventional recordings, on dense arrays, spikes are detectable on multiple channels. (cid:140)ese duplicate spikes pose two signi(cid:128)cant problems for traditional spike sorting algorithms. First, the amount of computation and memory used for processing each detected event increases with the number of duplicates. Second, the rate of misclassi(cid:128)cation in spike sorting potentially increases since each duplicate spike must be sorted into the same event. To avoid the pitfalls associated with duplicate spikes, it is suggested to identify and remove du- plicates during detection. One naive method for duplicate removal is to remove all but the largest amplitude spike in a radius that encompasses the spatial footprint of the event. (cid:140)is method will remove almost every duplicate event, but as the radius of duplicate removal increases, so does the number of spikes removed that are not associated with the original event. A more rigorous method for duplicate removal involves keeping the largest amplitude spikes and removing all spikes in a radius that have decayed in amplitude. (cid:140)is method allows for the separating of near-synchronous events that are in the same spatial area of the array. Its success, however, relies on the assump- tion that the timing of spikes from the same event on nearby electrodes is almost identical and only weakly in(cid:131)uenced by noise and that the signal spatially decays away from its current source (Hagen et al., 2015). 3.3 Feature Extraction (cid:140)e relevant signal a spike causes in extracellular recordings lasts around 3 ms, which, depending on the acquisition rate, may correspond to up to 90 data points per event. However, spike shapes are highly redundant and can be e(cid:129)ciently represented in a low dimensional space. (cid:140)us, an appropriate projection method can be used to compute a small number of features for each event, which can be more e(cid:129)ciently clustered than raw waveforms. (cid:140)e most common feature extraction method for extracellular spikes is Principal Component Analysis (PCA), performed on whitened and peak-aligned spike waveforms. PCA (cid:128)nds princi- pal components, or orthogonal basis vectors, whose directions maximize the variance in the data. Extracellular spikes can be summarized well by just 3-4 principal components, a manageable di- mensionality for most clustering algorithms. (Adamos et al., 2008). Other less frequently used methods include independent component analysis (ICA) (Hermle et al., 2004) and wavelet decom- position ((cid:139)iroga et al., 2004). A comparison of these methods showed that the performance of sorting algorithms depends not only on the feature extraction method employed, but also on the clustering algorithm ((cid:139)iroga et al., 2004). In practice, the comparably low computational cost and relative e(cid:130)ectiveness of PCA in discriminating between di(cid:130)erent neurons and neuron types makes it particularly suitable for large scale recordings (Adamos et al., 2008). To reduce memory load, the PCA decomposition can be evaluated for a subset of events from a large recording and all events can be projected along the chosen dimensions e(cid:129)ciently in batches (Hilgen et al., 2017). 3.4 Clustering spatio-temporal event footprints (cid:140)ere are (cid:128)ve fundamental problems with the clustering phase. (cid:140)e (cid:128)rst problem is that the ex- tracellular waveform of neurons are known to change amplitude and shape during bursting (Fee 6 et al., 1996). (cid:140)e second problem is that some recorded waveforms are distorted by overlapping action potentials from synchronous, spatially-local events. (cid:140)is occurs frequently in dense arrays and usually exist as outliers in the chosen feature space. (cid:140)e third problem is that electrodes can dri(cid:137) in the extracellular medium, changing the relative position of each neuron to the electrodes. Dri(cid:137) distorts waveform shapes over the duration of the recording. (cid:140)e fourth problem is that the duplication of spikes over neighboring channels can lead to refractory period violations or mis- classi(cid:128)cations. (cid:140)e (cid:128)(cid:137)h and (cid:128)nal problem is that the number of observed neurons is unknown, which requires the use of non-parametric clustering algorithms or requires the user to estimate the number of neurons for a parametric clustering algorithm. (cid:140)e choice of the clustering algorithm will be determined by the speed and scalability con- siderations, by hypotheses over the typical shape of a cluster in this space, and by how well the algorithm can deal with the previously listed problems. Many spike sorting methods cluster by (cid:128)(cid:138)ing Gaussian Mixture Models (GMMs), modelling the feature density pro(cid:128)les as a sum of Gaus- sians (Harris et al., 2000; Rossant et al., 2016), or by (cid:128)(cid:138)ing a mixture of t-distributions (Shoham et al., 2003). (cid:140)e unknown number of actual neurons can be introduced as a latent variable, and the inference problem be solved with the expectation-maximisation (EM) algorithm. Bayesian ap- proaches, which also quantify parameter uncertainty, have also been introduced (Wood and Black, 2008). (cid:140)ese approaches, however, only perform well for single channels and are conceptually and computationally hard to scale up to large, full-chip datasets. More recent clustering algorithms for spike sorting are density-based. Density-based algo- rithms generally detect peaks or high-density regions in the feature space that are separated by low-density regions. (cid:140)ese algorithms are non-parametric, allowing the classi(cid:128)cation pipeline to be fully automatic, however, the number of clusters found can depend heavily on both hyper- parameters and the chosen feature space. Density-based clustering algorithms have been imple- mented for spike sorting with promising results (Hilgen et al., 2017; Chung et al., 2017). For dense arrays, an added complication arises since the information contained in event foot- prints cannot be used directly for sorting spikes, since it is unknown which channels contain responses of a single neuron and how many neurons cause the observed responses. (cid:140)e resulting combinatorial explosion can be dealt with in three ways: Masked clustering A straight-forward way to reduce the dimensionality of the clustering prob- lem is to include only channels with detectable responses for each event. Classical expectation maximisation on a mixture model is then possible when the irrelevant parts of the data are masked out and replaced with a tractable noise model (Kadir et al., 2014). (cid:140)is strategy produces excellent results with the help of a semi-automated re(cid:128)nement step. (Rossant et al., 2016). A main limitation is, however, a super-linear scaling with the number of recordings channels, which makes it less suitable for the latest generations of large-scale arrays. Template matching Since the raw recorded signal can be linearly decomposed into a mixture of footprints from di(cid:130)erent neurons (Segev et al., 2004), template matching has been a successful strategy for spike sorting, and implementations are available that scale up to thousands of channels (Pachitariu et al., 2016; Lee et al., 2017; Yger et al., 2018). (cid:140)is approach has two steps. First a col- lection of spatio-temporal footprints is obtained in a single pass over the data and dimensionality reduction and clustering is used to build templates for single neurons. In a second pass, all events 7 are assigned to the most likely template or combination of templates in the case of temporally overlapping events. A major advantage of template matching that temporally overlapping spikes are naturally ac- counted for through addition of two relevant templates. (cid:140)is makes it very suitable for recordings with high (cid:128)ring rates and correlations between nearby neurons. A potential limitation is that neurons spiking at very low rates may remain undiscovered as no reliable template can be built through averaging. Moreover, current implementations require a (cid:128)nal manual curation step. (cid:140)is is, however, simpli(cid:128)ed by correcting the assignment based on templates, which can be merged or split, rather than based on single events. Spike localisation As explained above, the spatial spike footprint allows event localisation through an estimation of the barycentre from the peak event amplitudes in nearby channels. (cid:140)is produces density maps with clear, isolated peaks in event density, which represent spikes from single or multiple, nearby neurons (see Figure 2 for an example). A two-dimensional density map can be clustered very e(cid:129)ciently, and the combination of locations and waveform features obtained through dimensionality reduction allows successful separation of nearby neurons. Density-based clustering algorithms have been successfully employed to solve this task: DPClus, based on the identi(cid:128)cation of density peaks (Jun et al., 2017a), ISO-SPLIT to grow uni-modal clusters from small seeds (Chung et al., 2017) and Mean Shi(cid:137), which herds data points towards high-density areas (Hilgen et al., 2017). Of all methods discussed here, spike localisation and clustering potentially has the best com- putational performance, since the actual computation is performed on a data set with much lower dimensionality than the original data (Hilgen et al., 2017; Jun et al., 2017a). Because the number of dimensions in the clustering step has to be kept small, it also discards useful information. However, usually locations and spatio-temporal waveform features exhibit substantial redundancy (Hilgen et al., 2017), making this approach the most suitable for very large arrays. 4 Evaluation (cid:140)e evaluation of spike detection and sorting quality is complicated by data volume and com- plexity, which makes both manual and automated curation challenging. It is however possible to assess the quality of an algorithm using data with ground truth annotation. Moreover, methods for post-hoc quality assessment of desirable properties of single units can be used to accept or reject units found through spike sorting. Speci(cid:128)cally, the desired result of a spike sorting pipeline to minimise the false detection of noise as spikes (false positives in detection), and the number of real spikes le(cid:137) undetected (false negatives in detection). Moreover, it should not assign spikes to the wrong neuron, hence it should minimise false positives and negatives in a cluster assignment. When ground truth annotations are available, false positives and false negatives can be easily counted. A direct, but technically challenging method to obtain ground truth information, is the simultaneous recording of a single neuron, together with an array recording, which will then be analysed using the spike sorting algorithm in question. (cid:140)ree such data sets recorded with dense arrays are currently available, two from the rat cortex recorded in vivo (Neto et al., 2016; Marques- Smith et al., 2018), and one from the mouse retina recorded in vitro (Yger et al., 2018). In both cases, a single juxtacellular electrode placed very closely to the array reliably recorded all spikes from 8 a single neuron. A systematic analysis of spike sorting has shown a clear relationship between measured spike amplitude and classi(cid:128)cation accuracy with errors strongly increasing for events smaller than 50 mV (Yger et al., 2018). (cid:140)is important result can help motivate exclusion of units with weaker signals. Ground truth for spike sorting can also be produced by simulations. Recently, it has become possible to simulate a complete biophysical forward model for recorded extracellular potentials in neural tissue (Hagen et al., 2015). (cid:140)is has produced several data sets that are now used to benchmark spike sorting algorithms (see e.g. Lee et al. (2017)). In another study, ground truth data was generated by superimposing synthetic spikes onto a recording from an empty array. (cid:140)is data was used to evaluate the e(cid:130)ect of noise on event localisation accuracy and to discover that localisation is inevitably a trade-o(cid:130) between position uncertainty and bias (Muthmann et al., 2015). It is an open question, however, how well results collected from simulated data generalise, since the precise noise model, which may di(cid:130)er between recording systems, impacts spike sorting algorithm performance (Muthmann et al., 2015). Finally, for cases where no ground truth data is available, Hill et al. (2011) proposed a set of metrics that should accompany all spike sorting methods as an evaluation of their reliability. (cid:140)eir metrics, applied a posteriori, are based on di(cid:130)erent features of the sorted dataset, which can be summarised as follows: • (cid:140)e waveforms in each cluster. (cid:140)e average waveform can present non-biological features, hinting that the cluster may be a collection of wrongly detected events. Additionally, if properties of the typical waveform change over time, this may be a sign of neurons dri(cid:137)ing away from their initial position on the chip. Finally, anomalous variability of each feature above the noise level may be a sign that multiple neurons contributed to the same cluster. • (cid:140)e times of all spikes in a cluster. Violations of the refractory period show the cluster contains false positives: these can be studied via the autocorrelation function or inter-spike histogram of each cluster. • (cid:140)e amplitudes of action potentials. A sharp drop in the amplitude distribution, caused by the detection threshold, signi(cid:128)es that the la(cid:138)er has introduced an arti(cid:128)cial bias. • (cid:140)e separation between pairs of clusters. Ample, sharp interfaces between clusters mean the properties of each neuron's spikes overlap in the selected feature space. If this occurs, there will be a theoretical minimum of false positives and negatives due to the incorrect assignment of events to the wrong cluster. (cid:140)e last point can be evaluated by re-examining a group of clustered neurons with a mixture model (usually Gaussian), which can be (cid:128)t using more features than the original algorithm. Assuming that this (cid:128)t is at least as reliable as the original sorting algorithm, a comparison of the two assign- ments is informative regarding the reliability of each unit. A statistic summarizing all these tests can then be used to exclude events and units post hoc. Using this method, detection and clustering parameters do not have to be adjusted carefully prior to each analysis. 9 Name and reference Kilosort (Pachitariu et al. (2016)) github.com/cortex-lab/KiloSort YASS (Lee et al. (2017)) yass.readthedocs.io Herding Spikes (Hilgen et al. (2017)) github.com/mhhennig/HS2 MountainSort (Chung et al. (2017)) github.com/(cid:131)atironinstitute/mountainsort JRCLUST (Jun et al. (2017a)) jrclust.org SpyKING CIRCUS (Yger et al. (2018)) spyking-circus.rtfd.org Method Notes TM TM SL+D D SL+D TM detection template GPU support; MATLAB based; semi- automated (cid:128)nal curation. network-based Neural (GPU); outlier triaging; matching; clustering. Fast and scalable; tested on multiple array geometries Fully automatic; graphi- cal user interface; unique clustering method Probe dri(cid:137) correction; GPU support. scalable; GPU support; tested on many datasets; robust to overlapping spikes; graphical user interface. Table 1: Summary of the most recent spike sorting methods developed for large, dense arrays. For a summary of older algorithms -- mostly for smaller, sparser arrays -- see Bestel et al. (2012). TM = Template Matching; SL = Spike Localisation; D = Density-based clustering (see section 3.4) 5 Outlook In this chapter, we discussed the existing methodology for recovering single neuron activity from high density recordings and the challenges and problems that each approach faces. Six freely available spike sorting pipelines for large-scale extracellular arrays and the methods they use are summarised in Table 1. For more information on their unique advantages and disadvantages, please review their associated references. Since inaccurate detection and sorting can in(cid:131)uence subsequent analysis of neural populations (Ventura and Gerkin, 2012), manual curation steps are o(cid:137)en still required to guarantee good data quality. However, the recent methods we summarised in this chapter take signi(cid:128)cant steps in increasing the speed, automatisation, and accuracy of the spike sorting pipeline. Looking forward, it may be possible to apply novel machine learning techniques to improving spike sorting. (cid:140)is has already been put into practice with a recent spike sorting algorithm where a neural network is used to improve detection of neural events (Lee et al., 2017). Although neural networks are showing promising results in detection, it may be possible to (cid:128)nd new breakthroughs in both feature extraction and in classi(cid:128)cation using these methods. Moreover, a neural network approach may have the potential of encompassing all of the spike sorting steps within a single model. A challenge when using these machine learning algorithms, however, is the di(cid:129)culty of obtaining ground truth data, which is poorly available and usually under speci(cid:128)c experimental conditions that may not generalize to other data sets. Increased automation also means more work is needed in developing reliable methods for validation and quality control of spike sorting results. (cid:140)e introduction of synthetic (Hagen et al., 10 2015) and experimetal ground truth datasets (Neto et al., 2016; Yger et al., 2018) is an important step forward in this direction. A standardisation, both of the sorting pipeline and of its evaluation, should be considered among the next objectives of the spike sorting community. A joint e(cid:130)ort should be made in order to guarantee that methods are intuitive to use and results are easy to compare. References Adamos, D. A., Kosmidis, E. K., and (cid:140)eophilidis, G. (2008). Performance evaluation of pca-based spike sorting algorithms. Computer Methods and Programs in Biomedicine, 91(3):232 -- 244. Ballini, M., Muller, J., Livi, P., Chen, Y., Frey, U., Ste(cid:138)ler, A., Shadmani, A., Viswam, V., Jones, I. L., Jackel, D., Radivojevic, M., Lewandowska, M. K., Gong, W., Fiscella, M., Bakkum, D. J., Heer, F., and Hierlemann, A. (2014). A 1024-channel CMOS microelectrode array with 26,400 electrodes for recording and stimulation of electrogenic cells in vitro. IEEE Journal of Solid-State Circuits, 49(11):2705 -- 2719. Berdondini, L., van der Wal, P. D., Guenat, O., de Rooij, N. F., Koudelka-Hep, M., Seitz, P., Kauf- mann, R., Metzler, P., Blanc, N., and Rohr, S. (2005). High-density electrode array for imaging in vitro electrophysiological activity. Biosensors & Bioelectronics, 21(1):167 -- 74. Bestel, R., Daus, A. W., and (cid:140)ielemann, C. (2012). A novel automated spike sorting algorithm with adaptable feature extraction. Journal of Neuroscience Methods, 211(1):168 -- 178. Buzs´aki, G. and Mizuseki, K. (2014). (cid:140)e log-dynamic brain: how skewed distributions a(cid:130)ect network operations. Nature Reviews Neuroscience, 15(4):264. Chung, J. E., Magland, J. F., Barne(cid:138), A. H., Tolosa, V. M., Tooker, A. C., Lee, K. Y., Shah, K. G., Felix, S. H., Frank, L. M., and Greengard, L. F. (2017). A fully automated approach to spike sorting. Neuron, 95(6):1381 -- 1394. Dimitriadis, G., Neto, J. P., Aarts, A., Alexandru, A., Ballini, M., Ba(cid:138)aglia, F., Calcaterra, L., David, F., Fiath, R., Frazao, J., et al. (2018). Why not record from every channel with a cmos scanning probe? bioRxiv, page 275818. Eversmann, B., Jenkner, M., Hofmann, F., Paulus, C., Brederlow, R., Holzap(cid:131), B., Fromherz, P., Merz, M., Brenner, M., Schreiter, M., Gabl, R., Plehnert, K., Steinhauser, M., Eckstein, G., Schmi(cid:138)- landsiedel, D., and (cid:140)ewes, R. (2003). A 128 128 CMOS Biosensor Array for Extracellular Record- ing of Neural Activity. IEEE Journal of Solid-State Circuits, 38(12):2306 -- 2317. Fee, M. S., Mitra, P. P., and Kleinfeld, D. (1996). Variability of extracellular spike waveforms of cortical neurons. Journal of Neurophysiology, 76(6):3823 -- 3833. Frey, U., Sedivy, J., Heer, F., Pedron, R., Ballini, M., Mueller, J., Bakkum, D., Ha(cid:128)zovic, S., Faraci, F. D., Greve, F., Kirstein, K. U., and Hierlemann, A. (2010). Switch-matrix-based high-density microelectrode array in CMOS technology. IEEE Journal of Solid-State Circuits, 45(2):467 -- 482. 11 Hagen, E., Ness, T. V., Khosrowshahi, A., Sørensen, C., Fyhn, M., Ha(cid:137)ing, T., Franke, F., and Einevoll, G. T. (2015). ViSAPy: A Python tool for biophysics-based generation of virtual spiking activity for evaluation of spike-sorting algorithms. Journal of Neuroscience Methods, 245:182 -- 204. Harris, K. D., Henze, D. A., Csicsvari, J., Hirase, H., and Buzs´aki, G. (2000). Accuracy of tetrode spike separation as determined by simultaneous intracellular and extracellular measurements. Journal of Neurophysiololgy, 84(1):401 -- 414. Hermle, T., Schwarz, C., and Bogdan, M. (2004). Employing ica and som for spike sorting of multielectrode recordings from cns. Journal of Physiology-Paris, 98(4-6):349 -- 356. Hilgen, G., Sorbaro, M., Pirmoradian, S., Muthmann, J.-O., Kepiro, I. E., Ullo, S., Ramirez, C. J., Encinas, A. P., Maccione, A., Berdondini, L., et al. (2017). Unsupervised spike sorting for large- scale, high-density multielectrode arrays. Cell Reports, 18(10):2521 -- 2532. Hill, D. N., Mehta, S. B., and Kleinfeld, D. (2011). (cid:139)ality metrics to accompany spike sorting of extracellular signals. Journal of Neuroscience, 31(24):8699 -- 705. Hrom´adka, T., Deweese, M. R., and Zador, A. M. (2008). Sparse representation of sounds in the unanesthetized auditory cortex. PLoS Biology, 6(1):e16. Jun, J. J., Mitelut, C., Lai, C., Gratiy, S., Anastassiou, C., and Harris, T. D. (2017a). Real-time spike sorting platform for high-density extracellular probes with ground-truth validation and dri(cid:137) correction. bioRxiv, page 101030. Jun, J. J., Steinmetz, N. A., Siegle, J. H., Denman, D. J., Bauza, M., Barbarits, B., Lee, A. K., Anastas- siou, C. A., Andrei, A., Aydın, C¸., et al. (2017b). Fully integrated silicon probes for high-density recording of neural activity. Nature, 551(7679):232. Kadir, S. N., Goodman, D. F., and Harris, K. D. (2014). High-dimensional cluster analysis with the masked em algorithm. Neural Computation, 26(11):2379 -- 2394. Lee, J. H., Carlson, D. E., Razaghi, H. S., Yao, W., Goetz, G. A., Hagen, E., Ba(cid:138)y, E., Chichilnisky, E., Einevoll, G. T., and Paninski, L. (2017). Yass: Yet another spike sorter. In Advances in Neural Information Processing Systems, pages 4005 -- 4015. Lewicki, M. S. (1998). A review of methods for spike sorting: the detection and classi(cid:128)cation of neural action potentials. Network, 9(January):R53 -- R78. Lopez, C. M., Mitra, S., Putzeys, J., Raducanu, B., Ballini, M., Andrei, A., Severi, S., Welkenhuysen, M., Van Hoof, C., Musa, S., et al. (2016). 22.7 a 966-electrode neural probe with 384 con(cid:128)gurable channels in 0.13 µm soi cmos. In Solid-State Circuits Conference (ISSCC), 2016 IEEE International, pages 392 -- 393. IEEE. Marques-Smith, A., Neto, J. P., Lopes, G., Nogueira, J., Calcaterra, L., Frazo, J., Kim, D., Phillips, M. G., Dimitriadis, G., and Kamp(cid:130), A. R. (2018). Recording from the same neuron with high- density cmos probes and patch-clamp: a ground-truth dataset and an experiment in collabora- tion. bioRxiv, page 370080. 12 Muller, J., Ballini, M., Livi, P., Chen, Y., Radivojevic, M., Shadmani, A., Viswam, V., Jones, I. L., Fiscella, M., Diggelmann, R., Ste(cid:138)ler, A., Frey, U., Bakkum, D. J., Hierlemann, A., Muller, J., Ballini, M., Livi, P., Chen, Y., Radivojevic, M., Shadmani, A., Viswam, V., Jones, I. L., Fiscella, M., Diggelmann, R., Ste(cid:138)ler, A., Frey, U., Bakkum, D. J., and Hierlemann, A. (2015). High-resolution CMOS MEA platform to study neurons at subcellular, cellular, and network levels. Lab on a Chip, 15(13):2767 -- 2780. Muthmann, J.-O., Amin, H., Sernagor, E., Maccione, A., Panas, D., Berdondini, L., Bhalla, U. S., and Hennig, M. H. (2015). Spike Detection for Large Neural Populations Using High Density Multielectrode Arrays. Frontiers in Neuroinformatics, 9(December):1 -- 21. Neto, J. P., Lopes, G., Frazao, J., Nogueira, J., Lacerda, P., Baiao, P., Aarts, A., Andrei, A., Musa, S., Fortunato, E., et al. (2016). Validating silicon polytrodes with paired juxtacellular recordings: method and dataset. Journal of Neurophysiology, 116(2):892 -- 903. Obien, M. E. J., Deligkaris, K., Bullmann, T., Bakkum, D. J., and Frey, U. (2015). Revealing neuronal function through microelectrode array recordings. Frontiers in Neuroscience, 9(JAN):423. Pachitariu, M., Steinmetz, N. A., Kadir, S. N., Carandini, M., and Harris, K. D. (2016). Fast and accu- rate spike sorting of high-channel count probes with kilosort. In Advances in Neural Information Processing Systems, pages 4448 -- 4456. Panas, D., Amin, H., Maccione, A., Muthmann, O., van Rossum, M., Berdondini, L., and Hennig, M. H. (2015). Sloppiness in Spontaneously Active Neuronal Networks. Journal of Neuroscience, 35(22):8480 -- 8492. (cid:139)iroga, R. Q., Nadasdy, Z., and Ben-Shaul, Y. (2004). Unsupervised spike detection and sorting with wavelets and superparamagnetic clustering. Neural Computation, 16(8):1661 -- 87. Rey, H. G., Pedreira, C., and (cid:139)ian (cid:139)iroga, R. (2015). Past, present and future of spike sorting techniques. Brain Research Bulletin, 119:106 -- 117. Rossant, C., Kadir, S. N., Goodman, D. F. M., Schulman, J., Hunter, M. L. D., Saleem, A. B., Grosmark, A., Belluscio, M., Den(cid:128)eld, G. H., Ecker, A. S., Tolias, A. S., Solomon, S., Buzs´aki, G., Carandini, M., and Harris, K. D. (2016). Spike sorting for large, dense electrode arrays. Nature Neuroscience, 19(4):634 -- 641. Segev, R., Goodhouse, J., Puchalla, J., and Berry II, M. J. (2004). Recording spikes from a large fraction of the ganglion cells in a retinal patch. Nature Neuroscience, 7(10):1155. Shoham, S., Fellows, M. R., and Normann, R. A. (2003). Robust, automatic spike sorting using mixtures of multivariate t-distributions. Journal of Neuroscience Methods, 127(2):111 -- 122. Ventura, V. and Gerkin, R. C. (2012). Accurately estimating neuronal correlation requires a new spike-sorting paradigm. Proceedings of the National Academy of Sciences, 109(19):7230 -- 7235. Wood, F. and Black, M. J. (2008). A nonparametric bayesian alternative to spike sorting. Journal of Neuroscience Methods, 173(1):1 -- 12. 13 Yger, P., Spampinato, G. L., Esposito, E., Lefebvre, B., Deny, S., Gardella, C., Stimberg, M., Je(cid:138)er, F., Zeck, G., Picaud, S., et al. (2018). A spike sorting toolbox for up to thousands of electrodes validated with ground truth recordings in vitro and in vivo. eLife, 7:e34518. Yuan, X., Kim, S., Juyon, J., D'Urbino, M., Bullmann, T., Chen, Y., Ste(cid:138)ler, A., Hierlemann, A., and Frey, U. (2016). A microelectrode array with 8,640 electrodes enabling simultaneous full- frame readout at 6.5 kfps and 112-channel switch-matrix readout at 20 ks/s. In VLSI Circuits (VLSI-Circuits), 2016 IEEE Symposium on, pages 1 -- 2. IEEE. Zhang, J., Laiwalla, F., Kim, J. A., Urabe, H., Van Wagenen, R., Song, Y.-K., Connors, B. W., Zhang, Integrated device for optical stimulation and F., Deisseroth, K., and Nurmikko, A. V. (2009). spatiotemporal electrical recording of neural activity in light-sensitized brain tissue. Journal of neural engineering, 6(5):055007. 14
1412.1713
2
1412
2015-07-02T13:28:29
Networks that learn the precise timing of event sequences
[ "q-bio.NC" ]
Neuronal circuits can learn and replay firing patterns evoked by sequences of sensory stimuli. After training, a brief cue can trigger a spatiotemporal pattern of neural activity similar to that evoked by a learned stimulus sequence. Network models show that such sequence learning can occur through the shaping of feedforward excitatory connectivity via long term plasticity. Previous models describe how event order can be learned, but they typically do not explain how precise timing can be recalled. We propose a mechanism for learning both the order and precise timing of event sequences. In our recurrent network model, long term plasticity leads to the learning of the sequence, while short term facilitation enables temporally precise replay of events. Learned synaptic weights between populations determine the time necessary for one population to activate another. Long term plasticity adjusts these weights so that the trained event times are matched during playback. While we chose short term facilitation as a time-tracking process, we also demonstrate that other mechanisms, such as spike rate adaptation, can fulfill this role. We also analyze the impact of trial-to-trial variability, showing how observational errors as well as neuronal noise result in variability in learned event times. The dynamics of the playback process determine how stochasticity is inherited in learned sequence timings. Future experiments that characterize such variability can therefore shed light on the neural mechanisms of sequence learning.
q-bio.NC
q-bio
Noname manuscript No. (will be inserted by the editor) Networks that learn the precise timing of event sequences Alan Veliz-Cuba · Harel Z. Shouval · Kresimir Josi´c* · Zachary P. Kilpatrick* the date of receipt and acceptance should be inserted later Abstract Neuronal circuits can learn and replay firing patterns evoked by sequences of sensory stim- uli. After training, a brief cue can trigger a spa- tiotemporal pattern of neural activity similar to that evoked by a learned stimulus sequence. Net- work models show that such sequence learning can occur through the shaping of feedforward excita- tory connectivity via long term plasticity. Previous models describe how event order can be learned, but they typically do not explain how precise tim- ing can be recalled. We propose a mechanism for learning both the order and precise timing of event A. Veliz-Cuba Department of Mathematics, University of Houston, Houston TX 77204 USA E-mail: [email protected] H.Z. Shouval Department of Neurobiology and Anatomy, University of Texas Medical School, Houston TX 77030 USA E-mail: [email protected] K. Josi´c Department of Mathematics and Department of Biology, University of Houston, Houston TX 77204 USA E-mail: [email protected] Z.P. Kilpatrick Department of Mathematics, University of Houston, Houston TX 77204 USA E-mail: [email protected] *equal contribution sequences. In our recurrent network model, long term plasticity leads to the learning of the sequence, while short term facilitation enables temporally pre- cise replay of events. Learned synaptic weights be- tween populations determine the time necessary for one population to activate another. Long term plasticity adjusts these weights so that the trained event times are matched during playback. While we chose short term facilitation as a time-tracking pro- cess, we also demonstrate that other mechanisms, such as spike rate adaptation, can fulfill this role. We also analyze the impact of trial-to-trial vari- ability, showing how observational errors as well as neuronal noise result in variability in learned event times. The dynamics of the playback process determine how stochasticity is inherited in learned sequence timings. Future experiments that charac- terize such variability can therefore shed light on the neural mechanisms of sequence learning. Keywords serial recall · short term facilitation · long term plasticity 1 Introduction Networks of the brain are capable of precisely learn- ing and replaying sequences, accurately represent- ing the timing and order of the constituent events 5 1 0 2 l u J 2 ] . C N o i b - q [ 2 v 3 1 7 1 . 2 1 4 1 : v i X r a 2 Alan Veliz-Cuba et al. (Conway and Christiansen, 2001; Buhusi and Meck, 2005). Recordings in awake monkeys and rats re- veal neural mechanisms that underlie such sequence representation. After a training period consisting of the repeated presentation of a cue followed by a fixed sequence of stimuli, the cue alone can trig- ger a pattern of neural activity correlated with the activity pattern evoked by the stimulus sequence (Eagleman and Dragoi, 2012; Xu et al, 2012). Im- portantly, the temporal patterns of the stimulus- driven and cue-evoked activity are closely matched (Shuler and Bear, 2006; Gavornik and Bear, 2014). Sequence learning and replay has been iden- tified in a number of different brain areas. Re- cent electrophysiological recordings have located patterned neural activity in V1 (Xu et al, 2012; Gavornik and Bear, 2014) and V4 (Eagleman and Dragoi, 2012), corresponding to learned visual se- quences. Experiments on motor sequence learning found the underlying activity was coordinated by a combination of prefrontal, associative, and mo- tor cortical areas (Jenkins et al, 1994; Sakai et al, 1998). Training networks of the brain to replay mo- tor sequences is important since it allows quick mo- tor skill execution, faster than deliberate muscle control allows (Hikosaka et al, 2002). In addition, learning visual sequences can aid in experience- based prediction, so animals can react quickly to an unexpected chain of events (Meyer and Olson, 2011; Kok et al, 2012). Learning serial order is also an essential component of language and speech pro- duction in humans (Burgess and Hitch, 1999). In a similar way, music perception and production re- quires that humans learn to recognize and generate auditory-motor sequences (Zatorre et al, 2007). In total, sequence learning plays a large role in the daily cognitive tasks of a wide variety of animals. Various neural mechanisms have been proposed for learning the duration of a single event (Buono- mano, 2000; Rao and Sejnowski, 2001; Durstewitz, 2003; Reutimann et al, 2004; Karmarkar and Buono- mano, 2007; Gavornik et al, 2009), as well as the order of events in a sequence (Amari, 1972; Kle- infeld, 1986; Wang and Arbib, 1990; Abbott and Blum, 1996; Jun and Jin, 2007; Fiete et al, 2010; Brea et al, 2013). However, mechanisms for learn- ing the precise timing of multiple events in a se- quence remain largely unexplored. The activity of single neurons evolves on the timescale of tens of milliseconds. It is therefore likely that sequences on the timescale of seconds are represented in the activity of populations of cells. Recurrent network architecture could determine activity patterns that arise in the absence of input, but how this architec- ture can be reshaped by training to support pre- cisely timed sequence replay is not understood. Long term potentiation (LTP) and long term depression (LTD) are fundamental neural mecha- nisms that change the weight of connections be- tween neurons (Kandel, 2001). Learning in a wide variety of species, neuron types, and parts of the nervous system has been shown to occur through LTP and LTD (Alberini, 2009; Takeuchi et al, 2014; Nabavi et al, 2014). It is therefore natural to ask whether LTP and LTD can play a role in the learn- ing of sequence timing (Karmarkar and Buono- mano, 2007; Ivry and Schlerf, 2008; Bueti and Buono- mano, 2014), in addition to their proposed role in learning sequence order (Abbott and Blum, 1996; Fiete et al, 2010). We introduce a neural network model capa- ble of learning the timing of events in a sequence. The connectivity and dynamics in the network are shaped by two mechanisms: long term plasticity and short term facilitation. Long term synaptic plasticity allows the network to encode sequence and timing information in the synaptic weights, while slowly evolving short term facilitation can mark time during event playback. These ideas are quite general, and we show that they do not depend on the particulars of the time-tracking mechanisms we implemented. The impact of stimulus variabil- ity and neural noise is largely determined by the trajectory of the time-tracking process. Thus, we predict that errors in event sequence recall may be indicative of the mechanism that encodes them. Precise timing of event sequences 2 Material and Methods 2.1 Population rate model with short term facilitation Pyramidal cells in cortex form highly connected clusters (Song et al, 2005; Perin et al, 2011), which can correspond to neurons with similar stimulus tuning (Ko et al, 2011). We therefore considered a rate model describing the activity of N excita- tory populations (clusters) uj (j = 1, . . . , N ), and a single inhibitory population v. The excitatory and inhibitory populations were coupled via long range connections. Each population j received an exter- nal input Ij(t). Our model took the form: τ τf duj dt dpj dt τ dv dt = −uj + ϕ (Ij(t) + Ij,syn − θ) , = 1 − pj + (pmax − 1)uj, = −v + ϕ Zkuk − θv , (cid:33) (1) (cid:32) N(cid:88) k=1 N(cid:88) k(cid:54)=j where synaptic inputs to the jth population are given by Ij,syn = wjjuj + wjkpkuk − Lv. A complete description of the model functions and parameters is given in Table 1. Population firing rates ranged between a small positive value (the background firing rate), and a maximal value (the rate of a driven population), normalized to be 0 and 1, respectively. The baseline weight of the con- nection from population k to j was denoted by wjk. Connections within a population were denoted wjj, and these were not subject to short-term facilita- tion. Furthermore, the global inhibitory popula- tion, modeled by v, was typically active for very short epochs, so the effects of short term plasticity were not considered. These assumptions did not change our results, but made the analysis more transparent. Synapses between populations were subject to short term facilitation, and the facilitated connec- tion had "effective synaptic strength" wjkpk (Tsodyks 3 et al, 1998). Without loss of generality we assumed that short term facilitation varied between 1 and 2 so that the effective synaptic strength varied from wjk to 2wjk. Note that rescaling the maximal level of short term facilitation will simply rescale the relationship we will derive between the baseline synaptic weight and activation time of single pop- ulations. We assumed τf (cid:29) τ , in keeping with the observation that synaptic facilitation dynamics are much slower than changes in firing rates (Markram et al, 1998). In the absence of external stimulus or input from other populations, the dynamics of each pop- dt = −uj + ϕ(wjjuj − θ), ulation is described by τ duj and the stationary firing rate is given by the solu- tions of ϕ(wjjuj − θ) = uj. We typically took ϕ to be a Heaviside step function. Thus, if wjj > θ, there are two equilibrium firing rates: uj = 0 (in- active population) and uj = 1 (active population). This assumption simplified the analysis, however we show show in Section 4.2 that they are not es- sential. 2.2 Rate-based long term plasticity Connectivity between the populations in the net- work was subject to long term potentiation (LTP) and long term depression (LTD). Following experi- mental evidence (Bliss and Lømo, 1973; Dudek and Bear, 1992; Markram and Tsodyks, 1996; Sjostrom et al, 2001), connections were modulated using a rule based on pre- and post-synaptic activity with 'soft' bounds (Gerstner and Kistler, 2002). We made three main assumptions about the long term evolution of synaptic weight, w = wpre→post: (a) If the presynaptic population activity was low (upre ≈ 0), the change in synaptic weight was neg- ligible ( w(t) = 0); (b) If the presynaptic population was highly active (upre ≈ 1) and the postsynaptic population responded weakly (upost ≈ 0), then the synaptic weight decayed toward zero ( w(t) ∝ −w); and (c) If both populations had a high level of activity (upost ≈ 1 and upre ≈ 1), then synaptic weight increased towards an upper bound ( w(t) ∝ wmax − w). 4 Alan Veliz-Cuba et al. Table 1 Variables and parameters with their default values. The default values were used in all simulations, unless otherwise noted. References indicate work where similar parameter values were used or estimated: 1Graupner and Brunel (2012), 2Xu et al (2012), 3Gavornik and Bear (2014), 4Markram and Tsodyks (1996), 5Lundstrom (2015); 6Hausser and Roth (1997). Variables symbol Ij uj v pj wjk, w Tj, T symbol τ τf τw τa τs Tcue D D(cid:48) description external stimulus for excitatory population j non-dimensional firing rate of excitatory population j (maximum uj = 1) non-dimensional firing rate of global inhibitory population (maximum v = 1) level of facilitation of synapses from population j (baseline pj = 1) strength of excitation from population k to excitatory population j duration of stimulus Time parameters (default values in parenthesis) description timescale of neuronal firing (10ms 6) timescale of short term facilitation (1s 4) timescale of learning rule (150s 1) time scale of adaptation (400ms 5) time scale of synaptic inputs from other populations (50ms 6) duration of stimulus to trigger replay (50ms 2,3) delay in presynaptic firing affecting connections between populations (30ms 1) delay in presynaptic firing affecting connections within populations (20ms 1) Other parameters (default values in parenthesis) description firing rate response function (Heaviside step function) threshold for activation of excitatory population (0.5) threshold for activation of inhibitory population (0.5) maximum level of short term facilitation (2) strength of excitation from population k to inhibitory population (0.3) weight of global inhibition (0.6) strength of adaptation (1) learning rule threshold (1) maximum synaptic weight between populations (0.4852) maximum synaptic weight within populations (4.1312) minimum synaptic weight within populations (1.3488) strength of LTD between populations (150 1) strength of LTP between populations (3614.5 1) strength of LTD within populations (7500) strength of LTP within populations (267.86 1) symbol ϕ θ θv pmax Zk L b M wmax w(cid:48) max wmin γd γp γ(cid:48) γ(cid:48) d p Precise timing of event sequences 5 Similar assumptions have been used in previ- ous rate-based models of LTP/LTD (von der Mals- burg, 1973; Bienenstock et al, 1982; Oja, 1982; Miller, 1994), and it has been shown that calcium- based (Graupner and Brunel, 2012) and spike-time dependent (Clopath et al, 2010; Gjorgjieva et al, 2011) plasticity rules can be reduced to such rate- based rules (Pfister and Gerstner, 2006). Further- more, the fact that pre-synaptic activity is neces- sary to initiate either LTP or LTD is supported by experimental observations that plasticity de- pends on calcium influx through NMDA receptors (Malenka and Bear, 2004). A simple differential equation that implements these assumptions is τw dw dt = − γd w upre(t − D)(M − upost(t)) (2) + γp (wmax − w) upre(t − D) upost(t), where τw is the time scale, γd (γp) represents the strength of LTD (LTP), D is a delay accounting for the time it takes for the presynaptic firing rate to trigger plasticity processes, and M is a param- eter that determines the threshold and magnitude of LTD. We note that we initially model only the molecular processes that detect correlations in fir- ing rates, and thus set τw = 150s. We will later ex- tend this model to account for the longer timescales of synaptic weight changes (Section 4.4). Eq. (2) describes a Hebbian rate-based plas- ticity rule with soft bounds involving only linear and quadratic dependences of the pre- and post- synaptic rates (Gerstner and Kistler, 2002). Tem- poral asymmetry that accounts for the causal link between pre- and post-synaptic activity is incorpo- rated with a small delay in the dependence of pre- synaptic activity upre(t − D) (Gutig et al, 2003). This learning rule is a firing-rate version of the calcium-based plasticity model proposed by Graup- ner and Brunel (2012). 2.3 Encoding timing of event sequences Our training protocol was based on several recent experiments that explored cortical learning in re- sponse to sequences of visual stimuli (Xu et al, 2012; Eagleman and Dragoi, 2012; Gavornik and Bear, 2014). During a training trial, an external stimulus Ij(t) activated one population at a time. Each individual stimulus could have a different du- ration (Fig. 1B). We stimulated n populations, and enumerated them by order of stimulation; that is, population 1 was stimulated first, then population 2, and so on. This numbering is arbitrary, and the initial recurrent connections have no relation to this order. We denote the duration of input j by Tj. All inputs stop at Ttot = T1 + T2 + . . . + Tn. A sequence was presented m times. Repeated training of the network described by Eq. (1) with a fixed sequence drove the synap- tic weights wij to equilibrium values. We assumed that during sequence presentation, the amplitude of external stimuli Ij(t) was sufficiently strong to dominate the dynamics of the population rates, uj. Then, the activity of the populations during train- ing evolved according to: τ duj dt dwjk dt τw j = 1, . . . , n, = −uj + ϕ (Ij(t) − θ) , = −γdwjkuk(t − D)(M − uj(t)) (3) + γp(wmax − wjk)uk(t − D)uj(t), j (cid:54)= k. Thus, the timing of population activations mim- icked the timing of the input sequence, i.e. the training stimulus. 2.3.1 Synaptic weights for consecutive activations For simplicity, we begin by describing the case of two populations, N = 2, and we consider the thresh- old that determines the level of LTD equal to 1, M = 1, so that LTD is absent when the postsy- naptic population is active (upost = 1). Suppose that I1(t) = IS on [0, T1] and I2(t) = −IH , and I2(t) = IS on [T1, T1 + T2] and I1(t) = −IH , where IS and IH are large enough so that Eq. (3) is valid. The positive inputs with weight IS model feedfor- ward excitation to cells tuned to the cue from up- stream visual processing regions in thalamus. Neg- ative inputs with weight −IH model strong effec- tive feedforward inhibition to cortical cells that are 6 not tuned to the present cue (Wang et al, 2007; Haider et al, 2013). Representing the effect of feed- forward inhibition as static inputs simplified the model, and did not affect our results. We assumed that Ti > D, Ti (cid:29) τ , and τw (cid:29) τ , so that the stimulus was longer than the plastic- ity delay, and plasticity slower than the firing rate dynamics. Separation of timescales in Eq. (3) im- plies that uj ≈ ϕ(Ij(t) − θ), so the firing rate of populations 1 and 2 is approximated by u1(t) ≈ 1 on [0, T1] and zero elsewhere, and u2(t) ≈ 1 on [T1, T1 + T2] and zero elsewhere (Fig. 2A). Hence, during a training trial on a time interval [0, Ttot], we obtain from Eq. (3) the following piecewise equa- tion for the synaptic weight, w21, in terms of the duration of the first stimulus, T1, 0, − γd w21, τw γp(wmax − w21) τw t (cid:54)∈ [D, T1 + D], t ∈ [D, T1], , t ∈ [T1, T1 + D]. (4)  dw21 dt = Note that we assumed that u1(t) = 0 for t < 0. Eq. (4) allows the network to encode T1 using the weight w21. Namely, solving Eq. (4) we obtain w21(Ttot) =w21(0)e−T1γd/τw e−(γp−γd)D/τw + (1 − e−Dγp/τw )wmax, which relates the synaptic weight at the end of a presentation, w21(Ttot), to the synaptic weight at the beginning of the presentation, w21(0) (Fig. 2A). Thus, there is a recursive relation that relates the weight w21 at the end of the i + 1st stimulus to the weight at the end of the ith stimulus: wi+1 21 =wi 21e−T1γd/τw e−(γp−γd)D/τw + (1 − e−Dγp/τw )wmax. (5) As long as γp > γd, the sequence (wi to w∞ 21 := (1 − e−Dγp/τw )wmax 1 − e−T1γd/τw e−(γp−γd)D/τw , 21) converges (6) as seen in Fig. 2B. An equivalent expression also holds in the case of an arbitrary number of popu- lations. Alan Veliz-Cuba et al. The relative distance to the fixed point w∞ 21 is computed by noting that (for T1 fixed) wi+1 21 −w∞ 21/wi 21−w∞ 21 = e−T1γd/τw e−(γp−γd)D/τw , from which we calculate 21 ∝(cid:16) wi 21 − w∞ e−T1γd/τw e−(γp−γd)D/τw (cid:17)i . Thus, the sequence converges exponentially with the number of training trials, i. The relative dis- tance to the fixed point is proportional to e−mT1γd/τw , so the convergence is faster for larger values of T1, as shown in Fig. 2C. 2.3.2 Synaptic weights of populations that are not co-activated To compute the dynamics of w12, we note that dur- ing a training trial on the time interval [0, Ttot], the following piecewise equation governs the change in synaptic weight, dw12 dt = t (cid:54)∈ [T1 + D, T1 + T2 + D], 0, w12, t ∈ [T1 + D, T1 + T2 + D], − γd τw (cid:40) which can be solved explicitly to find w12(Ttot) = w12(0)e−T2γd/τw . (7) We can therefore write a recursive equation for the weight after the i + 1st stimulus in terms of the weight after the ith stimulus wi+1 12 = 0. Thus, w∞ 12e−T2γd/τw , 12 = wi which converges to w∞ jk = 0 for all pairs of populations (j, k) for which population j was not activated immediately after population k during training. In total, sequential activation of the populations leads to the strengthening only of the weights wj+1,j, while other weights are weak- ened. Precise timing of event sequences 7 2.4 Reactivation of trained networks To examine how a sequence of event timings could be encoded by our network, the first neural popu- lation in the sequence was activated with a short cue. Typically, this cue was of the form I1(t) = 1 for t ∈ [0, Tcue], I1(t) = 0 for t ∈ [Tcue,∞), and Ij(t) = 0 for j (cid:54)= 1 (Fig. 1C). During replay, aside from the initial cue, the activity in the network was generated through recurrent connectivity. We describe the case of two populations where the first population is cued, and remains active due to self-excitation (u1(t) ≈ 1), Fig. 3A. Since u2(0) = 0 and ϕ is the Heaviside step function, the equations governing the dynamics of the sec- ond population are = −u2 + ϕ(w21p1 − θ), du2 τ dt p1(t) = pmax + (1 − pmax)e−t/τf . (8) Thus, for population 2 to become active, w21p1(t) must have reached θ (Fig. 3A). The time T be- tween when u1 becomes active and u2 becomes active ("replay time") could be controlled by the synaptic weight w21. Fig. 3B shows the effect of changing the synaptic weight: For very small val- ues of the baseline weight w21, the effective synap- tic strength w21p1(t) never reaches θ and activation does not occur. Increasing the baseline weight w21 causes more rapid activation of the second popula- tion, and for very large weights w21 the activation is immediate. The weight required for a presynap- tic population to activate a postsynaptic popula- tion after T units of time is given in closed form by W(T ) := pmax + (1 − pmax)e−T /τf (9) θ . Similarly, the inverse of this function, T (w) := τf ln (cid:18) pmax − 1 (cid:19) , pmax − θ/w (10) gives the activation time as a function of the synap- tic weight (Fig. 3C). Note that Eq. (10) is valid for θ/pmax < w < θ. If w ≤ θ/pmax, then activation of the next population does not occur. If w ≥ θ, activation is immediate. To ensure the first population becomes inactive when population 2 becomes active, we assumed that global inhibition overcame the self excitation in the first population, w11 + w12p2(T )− L− θ < 0. Also, for the second population to remain active, we needed the self excitation plus the input re- ceived from population 1 to be stronger than the global inhibition; namely, w22 +w21p1(T )−L−θ = w22 − L > 0. These two inequalities are satisfied whenever w12 is small enough and L < wjj < L+θ. 2.5 Matching training parameters to reactivation parameters To guarantee that long term plasticity leads to a proper encoding of event times, it is necessary that the learned weight, w∞ 21 given by Eq. (6), matches the desired weight W(T ) given by Eq. (9). This can be achieved by equating the right hand sides of Eq. (6) and Eq. (9), so that (1 − e−Dγp/τw )wmax 1 − e−T γd/τw e−(γp−γd)D/τw = = θ/pmax 1 + e−T /τf (1 − pmax)/pmax . (11) Eq. (11) can be satisfied for all values of T by choosing parameters that satisfy τw/γd = τf , (1 − e−Dγp/τw )wmax = θ/pmax, e−(γp−γd)D/τw = (pmax − 1)/pmax. (12) Since there are fewer equations than model param- eters, there is a multi-dimensional manifold of pa- rameters for which Eq. (11) holds for all T . For instance, for fixed short-term facilitation parame- ters θ, pmax, and τf and restricting specific plastic- ity parameters τw and D, the appropriate γd, γp, and wmax can be determined using Eq. (12). This is how we determined the parameters in Figs. 4 and 5. The first relationship in Eq. (12) states τw/γd = τf , relating the timescale of short term facilitation 8 Alan Veliz-Cuba et al. to the timescale of long term plasticity through the depression amplitude parameter γd. It is important to note that this does not mean that the timescales of the two processes need to match. As stated in Table 1, following experimental data (Markram and Tsodyks, 1996; Alberini, 2009; Graupner and Brunel, 2012; Nabavi et al, 2014), we chose τf = 1s and τw = 150s for our simulations. This implies that γd = 150, for training to yield the correct weights. As we demonstrate in Supplementary Fig. 1, perturbing parameters of the long term plasticity process away from the optimal relationships deter- mined by Eq. (12) does alter the learned time. The relative size of errors depends on the parameters we perturb, and we find the model is most sensitive to perturbation of wmax. Perturbations of other pa- rameters such as γp and γd lead to to errors roughly equal in to the magnitude of the parameter pertur- bation (e.g., a 5% perturbation of γd leads to a 5% change in Treplay). For more detailed models, the analog of Eq. (12) is more cumbersome or impossible to obtain explic- itly. Specifically, when we incorporated noise into our models in the Section 4.1 and considered spike rate adaptation in the Section 4.3, we had to use an alternative approach. We found it was always possible to use numerical means to approximate parameter sets that allowed a correspondence be- tween the trained and desired weight for all possi- ble event times. A simple way of finding such pa- rameters was to use the method of least squares: We selected a range of stimulus durations, e.g. [.1s, 3s], and sampled timings from it, e.g. S = {.1s, .2s, .3s, . . . , 3s}. For each T ∈ S we computed the learned weight, wlearned(T, pars), where "pars" denotes the list of parameters to be determined. Then, we com- puted the replay time, Treplay(wlearned(T, pars)). We defined the "cost" function (cid:88) T∈S J(pars) = (Treplay(wlearned(T, pars)) − T )2 , and the desired parameters were given by parsbest := argminpars{J(pars)}. This approach was successful for different models and training protocols, and allowed us to find a (13) working set of parameters for models that included noise or different slow processes for tracking time. Eq. (13) can also be interpreted as a learning rule for the network parameters. Starting with arbi- trary network parameter values, any update mech- anism that decreases the cost function J will result in a network that can accurately replay learned se- quence times. 2.6 Training and replay simulations To test our model, we trained the network with a sequence of four events. Each event in the sequence corresponded to the activation of a single neuronal population (Fig. 4A). Since each population was inactivated (received strong negative input) when the subsequent populations became active, we also assumed that an additional, final population inac- tivated the population responding to the last event (additional population not shown in the figure). In- put during training was strong enough so that acti- vation of the different populations was only deter- mined by the external stimulus overriding global inhibition and recurrent excitation. During reactivation, the recurrent connections were assumed fixed. This assumption can be re- laxed if we assume that LTP/LTD are not imme- diate, but occur on long timescales, as in the Sec- tion 4.4. We used m = 10 training trials, the default pa- rameter values in Table 1, and estimated γd, γp, and wmax using Eq. (12). After the training tri- als were finished, we cued the first population in the sequence, using I1(t) = 1 for t ∈ [0, Tcue] and Ij(t) = 0 otherwise. We also started with this set of weights, and retrained the network with a novel sequence of stimuli. 3 Results 3.1 Training We explore sequence learning in a network model of neural populations, where each population is ac- tivated by a distinct stimulus or event. The initial Precise timing of event sequences 9 Fig. 1 The precise timing of training sequences is learned via long term plasticity. Each stimulus in the sequence is represented by a different color. A. Before training, network connectivity is random, and a cue does not trigger a sequential pattern of activity. B. During training, a sequence of events is presented repeatedly. Each event activates a corresponding neural population for some amount of time, which is fixed across presentations. Long term plasticity reshapes network architecture to encode the duration and order of these activations. C. After sufficient training, a cue triggers the pattern of activity evoked during the training period. Learned synaptic connectivity along with short term facilitation steer activity along the path carved by the training sequence (arrow width and contrast correspond to synaptic weight). connectivity between the populations is random (Fig. 1A). To make our analysis more transpar- ent, we initially consider a deterministic firing rate model, Eq. (1). Each individual neural population is bistable, having both a low activity state and a high activity state that is maintained through re- current excitation. Our results also hold for more biologically plausible firing rate response functions and are robust to noise (Sections 4.1 and 4.2). To train the network, we stimulated popula- tions in a fixed order, similar to the training paradigm used by Xu et al (2012); Eagleman and Dragoi (2012); Gavornik and Bear (2014). The duration of each event in the training sequence was arbi- trary (Fig. 1B), and each stimulus in the sequence drove a single neural population. Synaptic connec- tions between populations were plastic. To keep the model tractable, population activity was assumed to immediately impact the weight of synaptic con- nections. Our results also extend to a model with synaptic weights changing on longer timescales (Sec- tion 4.4). Changes in the network's synaptic weights de- pended on the firing rates of the pre- and post- synaptic populations (Bliss and Lømo, 1973; Bi- enenstock et al, 1982; Dudek and Bear, 1992; Markram and Tsodyks, 1996; Sjostrom et al, 2001). When a presynaptic population was active, either: (a) synapses were potentiated (LTP) if the postsynap- tic population was subsequently active or (b) syna- pses were depressed (LTD) if the postsynaptic pop- ulation was not activated soon after (Materials and Methods). Such rate-based plasticity rules can be derived from spike time dependent plasticity rules (Kempter et al, 1999; Pfister and Gerstner, 2006; Clopath et al, 2010). To demonstrate how the timing of events can be encoded in the network architecture, we start with timesequence trainingLTP LTDtime-codingrecurrent networkBCAtime (s)firingrate01timecuebefore trainingtime (s)firingrate01timecueafter trainingtime (s)firingrate011234arbitrary connectivitysequence presentationI1I2I3I412341234 10 Alan Veliz-Cuba et al. ods). While the first stimulus was present, popu- lation 1 was active and LTD dominated, decreas- ing the synaptic weight, w21, from population 1 to population 2. After T1 seconds, the first stimulus ended, and the second population was activated. However, population 1 did not become inactive in- stantaneously, and for some time both population 1 and 2 were active. During this overlap window, LTP dominated leading to an increase in synap- tic weight w21. Shortly after population 1 became inactive, changes in the weight w21 ceased, as plas- ticity only occurs when the presynaptic population is active. The initial and final synaptic weights (w0 21, respectively) can be computed in closed form (Materials and Methods). Repeated presentations of the training sequence leads to ex- ponential convergence of the synaptic weights, wi 21 (weight after ith training trial), to a fixed value (Fig. 2B). On the other hand, the synaptic weight w12 is weakened during each trial because the presy- naptic population 2 is always active after the post- synaptic population 1 (Materials and Methods). In the case of N populations, each weight wk+1,k will converge to a nonzero value associated with Tk, whereas all other weights will become negligible during replay. Thus, the network's structure even- tually encodes the order of the sequence. 21 and w1 The duration of activation in population 1, T1, determines the equilibrium value of the synaptic weight from population 1 to population 2, w∞ 21 (Ma- terials and Methods). For larger values of T1, LTD lasts longer, weakening w21 (Fig. 2C). Hence, weaker synapses are associated with longer event times. Reciprocally, weaker synapses lead to longer acti- vation times during replay (Section 3.2). As the stimulus duration, T1, determines the asymptotic synaptic weight, w∞ 21(T1), there is a map- ping T1 → w∞ 21(T1) from stimulus times to the re- sulting weights. Event timing is thus encoded in the asymptotic values of the synaptic weights. Fig. 2 Encoding timing in synaptic weights. A. Synap- tic connections evolve during training. When a presy- naptic population (1) is active and a postsynaptic pop- ulation (2) is inactive, LTD reduces the synaptic weight w21. When the populations (1 and 2) are co-active (over- lap window between dashed lines), LTP increases w21. Shortly after, global inhibition inactivates the presynap- tic population (1), so long term plasticity ceases (see Materials and Methods). As in the text, wi 21 denotes the weight at the end of the i-th trial. Arrow width and contrast correspond to synaptic weights. B. After sev- eral training trials, the synaptic weight wi 21 converges to a fixed point, w∞ 21, whose amplitude depends on the ac- tivation time of the presynaptic population. C. Starting from the same initial value, w0 21(T1) = 0.25, the weight 21(T1) converges to different values, w∞ wi 21(T1), depend- ing on the the training time, T1. Pink bar at T1=300ms corresponds to the value used in A and B. two populations (Fig. 2). During training, popula- tion 1 was stimulated for T1 seconds followed by stimulation of population 2 (Fig. 2A). The stimu- lus was strong enough to dominate the dynamics of the population responses (Materials and Meth- 3.2 A slow process allows precise temporal replay We next describe how the trained network replays sequences. The presence of a slow process, which 12001T1Aw012T10T1LTDLTPBT1firingratesynapticstrength w21timetrial 1trial 2trial 3trial 4sequencepresentationbeforeafterCduration of first stimulus (s)020.30.4timesynapticstrength w21synaptic strength w21T1timew1w0w1w∞w1w2w4w3w1(T1)w2(T1)w3(T1)w4(T1)w∞(T1)2121212121212121212121212121 Precise timing of event sequences 11 Such ramping models have previously been pro- posed as mechanisms for time-keeping(Buonomano, 2000; Durstewitz, 2003; Reutimann et al, 2004; Kar- markar and Buonomano, 2007; Gavornik et al, 2009). Without such a slow process, cued activity would result in a sequence replayed in the proper order, but information about event timing would be lost. For simplicity we focus on two populations, whe- re activity of the first population represents a timed event (Fig. 3). To simplify the analysis, we also assumed that synaptic weights are fixed during re- play. This assumption is not essential (Section 4.4). After population 1 is activated with a brief cue, it remains active due to recurrent excitation (Mate- rials and Methods). Meanwhile, short term facili- tation leads to an increase in the effective synaptic strength from population 1 to population 2. Popu- lation 2 becomes active when the input from pop- ulation 1 crosses an activation threshold (Fig. 3A). When both populations are simultaneously active, a sufficient amount of global inhibition is recruited to shut off the first population, which receives only weak input from population 2. The second popula- tion then remains active, as the strong excitatory input from the first population and recurrent exci- tation exceed the global inhibition. The weight of the connection from population 1 to population 2 determines how long it takes to extinguish the activity in the first population (Fig. 3B). This synaptic weight therefore encodes the time of this first and only event. We demon- strate how this principle extends to multiple event sequences in the Section 3.3. The time until the ac- tivation of the second population decreases as the initial synaptic weight increases, since a shorter time is needed for facilitation to drive the input from population 1 to the activation threshold (Fig. 3C). Note that when the baseline synaptic weight is too weak, synaptic facilitation saturates before the effective weight reaches the activation threshold, and the subsequent population is never activated. When the baseline synaptic weight is above the activation threshold, the subsequent population is activated instantaneously. Fig. 3 Replay timing. A. Once population 1 is acti- vated, the weight of the connection to population 2 slowly increases, eventually becoming strong enough to activate this next population in the sequence. B. Ac- tivation time T := T1 decreases monotonically with the weight of the connection between the populations, w := w21. For weak connections (wA = 0.33) the synapses must be strongly facilitated to reach the thresh- old. If the weight is larger (wB = 0.42) threshold is reached more quickly. For weights above the threshold (wC = 0.58 > θ), population 2 is activated immediately. Activation will not occur when the synaptic connection is smaller than θ/pmax. C. Activation time T1 plotted against the initial synaptic weight, w21. For intermedi- ate values of w21, the relationship is given by Eq. (10). Here wA, wB, wC and T A, T B are the same as in panel B. we assumed here to be short term facilitation, is critical. This slow process tracks time by ramping up until reaching a pre-determined threshold. An event's duration corresponds to the amount of time it takes the slow variable to reach this threshold. 12001synapseTT0Aθw21p1(t)w21p1(t)B12θimmediateactivationtimedactivationnoactivationfacilitatedsynapseθpmaxeffectivestrengthw21p1firingratetimetimeactivationtime0Cno activationwAwBimmediate activationTAactivation time (s)TBsynaptic strength1activation time (s)synaptic strengthwCθpmaxθwAwBwCTBTAvglobalinhibition 12 Alan Veliz-Cuba et al. Therefore, long term plasticity allows for en- coding an event time in the weight of the connec- tions between populations in the network, while short term facilitation is crucial for replaying the events with the correct timing. The time of acti- vation during cued replay will match the timing in the training sequence as long as training drives the synaptic weights to the value that corresponds to the appropriate event time (Fig. 2C): W(T ) := θ pmax + (1 − pmax)e−T /τf . (14) Here pmax and τf are the strength and timescale of facilitation (Materials and Methods). An exact match can be obtained by tuning parameters of the long term plasticity process so the learned weight matches Eq. (14). There is a wide range of param- eters for which the match occurs (Materials and Methods). We next show that the timing and or- der of sequences containing multiple events can be learned in a similar way. 3.3 Repeated presentation of the same sequence produces a time-coding feedforward network To demonstrate that the mechanism we discussed extends easily to arbitrary sequences, we consider a concrete sequence of four stimuli. We set the pa- rameters of the model so that the training parame- ters match the reactivation parameters (Eqs. 11-13 in Materials and Methods). We trained the network using the event sequence 1-2-3-4 (Fig. 4A), repeatedly stimulating the corre- sponding populations in succession. The duration of each population activation was fixed across tri- als. After each training trial, we cued the network by stimulating the first population for a short pe- riod of time to trigger replay (Materials and Meth- ods). Thus, after population 1 is activated the sub- sequent activity is governed by the network's ar- chitecture. As our theory suggests, the cue-evoked network activity pattern converged with training to the stimulus-driven activity pattern (Fig. 4A). Fig. 4 Learning and replaying event sequences. A. As the number of training trials increases, a cue results in an activation pattern that approaches that evoked by the training sequence. Network architecture is reshaped to encode the precise duration and order of events in the sequence, with stronger feedforward connections corre- sponding to shorter events (thicker and darker arrows correspond to larger synaptic weights). All weights are learned independently and training most strongly affects the weights wi→i+1. The event times were T1 = 0.6, T2 = 0.4, T3 = 1, and T4 = 0.5 seconds, with indices denoting the population stimulated. The initial weights ij = 0.025 for i (cid:54)= j and wii = 1. B. The network were w0 in the last row of A was retrained with the sequence T1 = 0.4, T4 = 1, T3 = 0.6, and T2 = 0.8, presented in the order 1-4-3-2. After 10 training trials, the cued network replays the new training sequence. We further tested whether the same network can be retrained to encode a sequence with a dif- ferent order of activation (1-4-3-2) with different event times. Fig. 4B shows that after training, the network encodes and replays the new training se- quence. Thus, the network architecture can be shaped by long term plasticity to encode an arbitrary se- training sequencetime (s)firing rateA01120101010time (s)01210Bcued activitynew training sequencetrial 1trial 2trial 3trial 101234cued activity (trial 10)12341234123412341432trained networkfiring rateretrained networkrearranged bystimulus order Precise timing of event sequences 13 Fig. 5 Effect of noise on learning and replay. The ef- fects of adding normally distributed noise to A the ob- served training time T := T1 (duration of first stimulus) and B neural activity uj. Starting from a uniform dis- tribution (p(w0 21)) and considering either a noise level of C σ = 0.1(cid:104)T(cid:105) in the training time or D σ = 0.03 in neural activity, the probability density function p(wi 21) converges to the steady state distribution p(w∞ 21) (this is nearly identical to w10 21). Increasing noise in the training times (E) or neural activity (F) widens the steady state distribution p(w∞ 21). The mean (dashed lines) and stan- dard deviation (shaded region) of w∞ 21(T ) are pictured in panel G when noise with standard deviation σ = 0.20(cid:104)T(cid:105) is added to the training times and H when noise with standard deviation σ = 0.05 is added to neural activity. As in Fig. 2C, the mean learned weight E [w∞ 21] decreases with (cid:104)T(cid:105). The mean of the replay time (dashed blue line) and its standard deviation (shaded cyan region) are plot- ted against (cid:104)T(cid:105) for the case of I noise in training times and J noise in neural populations. As the mean train- ing time (cid:104)T(cid:105) increases, so does the effect of noise on replay time, Treplay (grey lines show the diagonal line Treplay = T ). For a suitable choice of "corrected-for- noise" parameters, the effect of noise on the mean re- play time can be removed (dashed green line and green shaded region). Insets in I and J show the root-mean- square error in replay time as a function of training time for the parameters obtained from the deterministic case (blue) and corrected-for-noise parameters (green). we explored how noise impacts the training and recall of event sequences in our model. We exam- ined the effects of stochasticity in event times as well as noise in the network activity and the im- pact such variability has on the training and recall of event sequences. For simplicity, we focus on the case of two populations, but our results extend to sequences of arbitrary length. To examine the impact of variability in stimu- lus durations, we sampled T := T1 from a normal distribution (Fig. 5A) with mean (cid:104)T(cid:105) = 0.5 and variance σ2 = (cv (cid:104)T(cid:105))2, where cv = 0.1 is the co- efficient of variation. Randomness in the observed event time may be due to variability in the exter- nal world, temporal limitations on sensation (Butts et al, 2007), or other observational errors (Ma et al, 2006). We selected w0 := w0 21 from a uniform dis- tribution on [θ/pmax, θ]. Fig. 5C shows the evolu- tion of the probability density function of wm as m increases (using 20,000 initial w0's). The synap- Fig. 5 Effect of noise on learning and replay (caption continued on next page). quence of event times, and a brief cue evokes the replay of the learned sequence. 4 Extensions 4.1 Effects of noise To examine the impact of the many sources of vari- ability in the nervous system (Faisal et al, 2008), w10w2w0σ=0.15⟨T⟩σ=0.10⟨T⟩σ=0.20⟨T⟩θanalyticssimulations000200ww∞Tsynapticstrengthsynaptic strengthθpmaxθθpmaxθ02trained time (s)T02trainedtime (s)percentagepercentageθpmaxw∞synaptic strengthreplay time (s)02θθpmaxw∞synaptic strengthσ=0.05σ=0.03σ=0.06replay time (s)Treplay10101010percentagepercentagew10w2w0TreplaysimulationsCEGIDFHJ0202ww∞synapticstrengthsynaptic strengthθpmaxθθpmaxθ020.5RMSE⟨T⟩⟨T⟩trained time (s)⟨T⟩trained time (s)0201RMSETN(T,σ2)trial 1trial 2trial 3001TfiringratetimeABnoise in training timenoise in firing activity~T1T2T30202.1.2.2.4⟨T⟩TCVCV 14 Alan Veliz-Cuba et al. tic weight after the ith training, wi 21, is described by a probability density function that converges in the limit of many training trials. The peak (mode) of this distribution is the most likely value of the learned synaptic weight after repeated presenta- tion of the sequence (Fig. 5C). The variance of the learned synaptic weight, w∞ 21, increases mono- tonically with the variance in the training time, σ2 (Fig. 5E). We explored the effects of different levels of noise, σ = 0.1(cid:104)T(cid:105), 0.15(cid:104)T(cid:105), 0.2(cid:104)T(cid:105) (20,000 initial conditions for each), and estimated the probability density function of w∞ numerically (Fig. 5E). To see the effect of noise for different mean training durations, we estimated the mean and standard deviation of w∞ for (cid:104)T(cid:105) ranging from 0.1s to 2s (step size of 0.05s) using σ = 0.2(cid:104)T(cid:105) (5,000 initial conditions for each) (Fig. 5G). For a distribution of training times with mean (cid:104)T1(cid:105) we obtain a uni- modal probability density for the weights, p(w21). As in the noise-free case, the mode of this weight distribution decreases with (cid:104)T1(cid:105). Note that the pa- rameters given by Eq. (12), which guarantee that training and replay time coincide in the determin- istic case, may not be the same as the parameters needed when noise is present. We numerically esti- mated these parameters using Eq. (13) so the mean training time and mean replay time coincided. To determine how noise affects activation tim- ing during sequence replay, we compared the mean event time with the mean replayed time. Since the network parameters used here are those obtained from the noise-free case, we expect that replay times are biased. Indeed, Fig. 5I shows that activity dur- ing replay is slightly longer on average than the corresponding training event. Also, the variance in activation during replay increases with the mean duration of the trained event. We can search nu- merically and find a family of parameters for which the mean activation time during replay and train- ing coincide (Fig. 5I). Error and the coefficient of variation (CV) in replay time increases with the duration of the trained time (Fig. 5I, shaded re- gion and inset). We also estimated the mean learned synaptic weight and its variance analytically: Since the synap- tic weight, wi 21, evolves according to the rule 21A(T i) + C, wi+1 21 = wi where A(T ) := e−T γd/τw e−(γp−γd)D/τw and C := (1 − e−Dγp/τw )wmax, we obtain E[wi+1 21A(T i)] + C. 21 ] = E[wi Since wi lows that wi variables; then, 21 only depends on T j for j < i, it fol- 21 and A(T i) are independent random 21 ] = E[wi 21]E[A(T i)] + C, E[wi+1 and by taking the limit i → ∞ and solving for limi→∞ E[wi 21] we find µw := lim i→∞ E[wi 21] = C 1 − µA , where µA := E[A(T i)]. Squaring wi+1 C gives 21 = wi 21A(T i)+ (wi+1)2 = (wi)2A2(T i) + C 2 + 2wi 21A(T i)C, and then it similarly follows that (cid:0)E[(wi 21)2] − E[wi C 2σ2 A 21]2(cid:1) σ2 w := lim i→∞ , = A − µ2 A) (1 − µA)2(1 − σ2 A := E[A(T i)2] − E[A(T i)]2. To quantify where σ2 the average error, we computed numerically the mean and the standard deviation of the replay time (Fig. 5I). The root-mean-square error (RMSE) was computed by RMSE((cid:104)T(cid:105)) := E[((cid:104)T(cid:105) − Treplay)2], (cid:113) where the expected value is taken over the replay time, Treplay. The coefficient of variation (CV) was computed by (cid:112)V ar(Treplay) (cid:104)Treplay(cid:105) . CV ((cid:104)T(cid:105)) := To introduce neural noise (Fig. 5B), we added white noise to the rate equations of the populations during training and replay so that Precise timing of event sequences 15 duj = 1 τ [−uj + ϕ (Ij(t) + Ij,syn − θ)] dt + σdξj, where dξj is a standard white noise process with variance σ2. The analysis of the effect of noise in population activity is similar to the analysis per- formed on stimulus duration noise, the only dif- ference being that the noise level was σ = 0.03 in panel Fig. 5D, σ = 0.03, 0.05, 0.06 in panel Fig. 5F, and σ = 0.05 in panels Fig. 5H and 5J. After re- peated presentation of a sequence, the distribution of the learned synaptic weights converged (Fig. 5D). The variance of the synaptic weight increased mono- tonically with the variance of the noise (Fig. 5F), and the mean weight decreased monotonically with the event time (Fig. 5H). Since we used the pa- rameters found from the noise-free case, we ex- pect some bias in replay time. After training, the replayed event times are shorter than the corre- sponding events in the training sequence (Fig. 5J), and the effect is much more significant than the lengthening of times due to observation noise. This systematic bias in the replayed time error is due to the saturating nature of the time-tracking pro- cess, short term facilitation (Markram et al, 1998). Input to the second population remains close to threshold for longer periods of time for longer trained times, leading to more frequent noise-induced thresh- old crossings (Gardiner, 2004). However, parame- ters for which mean event time and mean replay time coincide can be found numerically (Materi- als and Methods). Interestingly, the CV attains a minimum in the vicinity of the timescale of short term facilitation (τf = 1s), suggesting the network best encodes events on the timescale of the slow process (Fig. 5J, lower inset). This principle holds across a range of short term facilitation timescales τf (Supplementary Material, Fig. 2). 4.2 Alternative firing rate response function ϕγ(x) = We next show that the mechanism for learning the precise timing of an event sequence does not de- pend on the particulars of the model. In previous where the parameter γ determines the steepness. Note that in the absence of other population in- Fig. 6 Alternative firing rate function. The mechanism for sequence learning and replay also works for other firing rate functions (see Materials and Methods). A. Using a nonlinear and nonsaturating response function ϕγ (u) = γu, long term plasticity still results in the coding of a training sequence as synaptic weights. B. After training, the cued network replays the training sequence similarly to the replay seen in Fig. 4A. √ sections, we used a Heaviside step function as the firing rate function and chose short term facilita- tion as the slow, time-tracking process. However, the principles we have identified do not depend on these specific choices. More general circuit models of slowly ramping units can learn and replay timed event sequences. The elements needed to learn and replay precise time durations are a chain of slow accumulators and a learning rule which modifies weights to precisely time a threshold-crossing event for each accumulator (Supplementary Material, Fig. 3). Precise replay relies on a threshold-crossing pro- cess which occurs as long as each population is bistable, having only low and high activity states rather than graded activity. Indeed, detailed spik- ing models (Litwin-Kumar and Doiron, 2012) and experimental recordings (Major and Tank, 2004) suggest that cell assemblies can exhibit multiple stable states. To test whether our conclusions hold with different firing rate response functions, we re- place the Heaviside function with the nonsaturat- ing function (Fourcaud-Trocm´e et al, 2003) (Fig. 6A) (cid:40) 0 √ if x < 0, if x ≥ 0, γx time (s)01210training sequencecued activity (trial 10)firing rate0nonlinear responsefunctionpre.post.φAB 16 puts, τ duj dt = −uj + ϕγ (wjjuj − θ) , which has steady states determined by the equa- tion u = ϕγ (wjju − θ). One of the stable steady states is u = 0 and there is a positive stable steady state, u∗, which is the largest root of the quadratic equation u2 − γwjju + γθ = 0. For simplicity, we normalize γ and self-excitation so that the stable states are u = 0 and u = 1; namely, we consider γ(cid:48) := γ/u2∗ and w(cid:48) jj := wjju∗. Since a population is activated when its input reaches the threshold due to short term facilita- tion, the derivations that led to Eqs. 9 and 10 are still valid for this model. However, the activation of a neuronal population (uj → 1) was delayed since ϕγ has finite slope. This delay was negligible when firing rate response was modeled by a Heavi- side function, and activation was instantaneous. To take this delay into account, we can modify Eqs. 9 and 10 to obtain W(T ) := pmax + (1 − pmax)e−(T−d)/τf θ , and T (w) := d + τf ln (cid:18) pmax − 1 pmax − θ/w (cid:19) , where d is a heuristic correction parameter to ac- count for the time it takes for uj to approach 1. Following the arguments that led to Eq. (12), we were able to derive constraints on parameters to ensure the correct timings are learned: 1 − e (M−1)γd+γp −D( 1 + (M−1)γd τw γp = τf , τw M γd  wmax = θ/pmax, ) (15) e(γd−γp)D/τw = pmax − 1 pmax ed/τf . For simulations we used the parameters M = 1.5, d = 30ms, γ = 3, and estimated γd, γp, and wmax using Eq. (15) (we then normalized ϕγ and wjj to Alan Veliz-Cuba et al. make 0 and 1 the stable firing rates). As in the pre- vious simulations, the number of presentations was m = 10; the durations of the events were 0.6s, 0.4s, 1s, and 0.5s for events 1, 2, 3 and 4, respectively. Network architecture converges, and the replayed activity matches the order and timing of the train- ing sequence (Fig. 6B). 4.3 An alternative slow process: spike rate adaptation We also examined whether spike frequency adapta- tion, i.e. a slow decrease in firing rates in response to a fixed input to a neural population, can play the role of a slow, time tracking process (Benda and Herz, 2003), instead of short term facilitation. In contrast to the case of short term facilitation, adaptation causes the effective input from one pop- ulation to decrease over time. In this case population activity was modeled by τ τa τs duj dt daj dt dsj dt τ dv dt = −uj + ϕ(wjjuj + sj − θ − Lv − aj), = −aj + buj, = −sj + wjkuk, = −v + ϕ Zkuk − θv N(cid:88) (cid:32) N(cid:88) k(cid:54)=j k=1 (cid:33) , where aj denotes the adaptation level of popula- tion j, τa is the time scale of adaptation, and b is the adaptation strength. Feedback between pop- ulations was assumed to be slower than feedback within a population; thus, the total input for pop- ulation j was split into self-excitation (wjjuj), and synaptic inputs from other populations (sj) which evolved on the time scale τs. Note that in the limit τs → 0, synapses are instantaneous. For a suitable choice of parameters, global in- hibition tracks activity faster than excitation be- tween populations. Then, when a population be- comes inactive due to adaptation, the level of global Precise timing of event sequences 17 Fig. 7 Alternative slow process based on spike fre- quency adaptation. A. When a population is active, LTP increases the synaptic weight w11. After becoming inactive, LTD decreases w11 (see Materials and Meth- ods). B. Synaptic weight wi 11 is updated after the ith training trial. After several trials, wi 11 converges to a fixed point w∞ 11 that depends on the activation time of the first population. C. Once population 1 is ac- tive, adaptation builds up until overcoming self excita- tion. This will occur when the effective strength, w11 − a1(t) ([excitation]−[adaptation]), crosses below θ + L ([threshold]+[inhibition]). D. Activation time T := T1 increases with synaptic weight w := w11. For strong self excitation (wA) adaptation takes longer to shut off the first population, so the next population in the sequence is activated later. Weaker self excitation (wB) will result in quicker extinction of activity in the first population, result in the next population activating sooner. For self excitation below θ + L ([threshold] + [inhibition]), the first population will inactivate immediately, resulting in immediate activation of the next population. If self ex- citation of the first population is greater than θ + L + b ([threshold] + [inhibition] + [maximum adaptation]), it will remain active indefinitely, and the subsequent pop- ulation is never activated. E. When the parameters of the long term plasticity process and the replay process match, the network can learn the precise timing of se- quences. analogous to wjk with the additional assumption that since wjj represented the synaptic weight within a population, it could not decrease below a cer- tain value wmin. Also, the parameters for long term plasticity within a population are allowed to be dif- ferent from the parameters for long term plasticity between populations. The learning rule was then τw dwjj dt = − γ(cid:48) − γ(cid:48) d(wjj − wmin)uj(t − D(cid:48))(1 − uj(t)) p(wjj − w(cid:48) max)uj(t − D(cid:48))uj(t). When the population was activated (u1(t) ≈ 1) for t ∈ [0, T1] (Fig. 7A), the changes in the weight w11 were governed by the piecewise differ- ential equation Fig. 7 Alternative slow process based on spike fre- quency adaptation (caption continued on next page). inhibition decreases, allowing subsequent popula- tions to become active. This means the weight of self excitation can encode timing. Thus, in this setup we modeled long term plasticity within a population as well. The learning rule for wjj was dw11 dt =  t (cid:54)∈ [D(cid:48), T1 + D(cid:48)] 0, max − w11) t ∈ [D(cid:48), T1] (w(cid:48) (w11 − wmin) t ∈ [T1, T1 + D(cid:48)]. γ(cid:48) p τw − γ(cid:48) d τw time (s)01210training sequencecued activity (trial 10)firing rateθ+Lw11-a1(t)12001T1A12T10T1LTDLTPBT1firingratesynapticstrength w11timetrial 1trial 2trial 3trial 4sequencepresentationbeforeaftertimesynapticstrength w11timew011w111w111w011w∞11w111w211w311w411w11-a1(t)C12001synapseTT0Deffectivestrengthfiringratetimetimeactivationtime0no activationimmediate activationTAactivation time (s)TBsynaptic strength1wAwBwCvw11-a1(t)Eθ+Lθ+L+b 18 Alan Veliz-Cuba et al. The following equation relates the synaptic weight at the end of a presentation, w11(Ttot), to the synap- tic weight at the beginning of the presentation, w11(0): w11(Ttot) =w11(0)e−T1γ(cid:48) maxe−D(cid:48)γ(cid:48) p/τw e(γ(cid:48) d/τw (1 − e−(T1−D(cid:48))γ(cid:48) d)D(cid:48)/τw p/τw ) p−γ(cid:48) + w(cid:48) + wmin(1 − e−D(cid:48)γ(cid:48) d/τw ). This recurrence relation between the weight at the i + 1st stimulus, wi+1 11 , and the weight at the ith stimulus, wi 11 converges to the limit (Fig. 7B) 11, implies that wi w∞ 11 = w(cid:48) maxe−D(cid:48)γ(cid:48) d/τw (1 − e−(T1−D(cid:48))γ(cid:48) d)D(cid:48)/τw 1 − e−T1γ(cid:48) p/τw e(γ(cid:48) wmin(1 − e−D(cid:48)γ(cid:48) p−γ(cid:48) p−γ(cid:48) d/τw ) d)D(cid:48)/τw p/τw e(γ(cid:48) 1 − e−T1γ(cid:48) + p/τw ) . Thus, for each stimulus duration a unique synaptic weight is learned. Also, as shown in previous sec- tions, w21 will converge to a fixed value and w12 is weakened. Note that in this case timing will be encoded as the weight of self-excitation, and the order will be encoded as the weights between pop- ulations. da1 dt du1 dt 1, and we obtain τ During replay a cue activates population 1, u1 = = −u1 + ϕ(w11− θ− a− L), = −a + b. Population 1 will become inac- τa tive (u1 ≈ 0) when w11 − a1(t) decreases to θ + L. Then, the next population will become active due to the decrease in global inhibition and the remain- ing feedback from the first population due to the slower dynamics of feedback between populations (Fig. 7C). The precise time of activation can be controlled by tuning the synaptic weight w11. Furthermore, since the activation time satisfies w11 − a1(T ) = θ + L, we have a formula that relates the synaptic weight to the activation time (Fig. 7D) w11 = W(T ) := θ + L + b(1 − e−T /τa ). When self excitation is too strong, adaptation will not affect the activity of the first population and deactivation will never occur. On the other hand, when self excitation is too weak, activation is not sustained and the population will be shut off imme- diately. To guarantee correct time coding and de- 11(T ) and W(T ) had to be approximately coding, w∞ equal for all T . The appropriate parameters could not be found in closed form, so we again resorted to finding them numerically using Eq. (13). For simulations we used the parameters Zk = 0.6, L = 0.8, γp = 3750, γd = 100, γ(cid:48) p = 267.86, γ(cid:48) d = 7500, wmax = 1.5, w(cid:48) max = 4.1312, and wmin = 1.3488. As in the previous simulations, the number of presentations was m = 10; the duration of the events were 0.6s, 0.4s, 1s, and 0.5s for events 1, 2, 3 and 4, respectively. This idea generalizes to any number of events and populations. Timing is encoded in the weight of the excitatory self-connections within a popula- tion, while sequence order is encoded in the weight of the connections between populations. Moreover, for a range of network parameters, the duration of the sequences during training and reactivation coincide (Materials and Methods). Presenting the event sequence used in Fig. 4A, the network can learn the precise timing and order of the events (Fig. 7E). 4.4 Incorporating long timescale plasticity Thus far, we have assumed that during sequence replay, synaptic connections remained unchanged (Section 3.3). However, if synaptic changes occur on the same timescale as the network's dynamics, and are allowed to act during replay, the network's architecture can become unstable. This problem can be solved by assuming that synaptic weights change slowly compared to network dynamics (Al- berini, 2009). We therefore extended our model so that long term plasticity occurs on more realistic timescales. The impact of rate covariation on the network's synaptic weights was modeled by a two step pro- cess: (a) rate correlation detection, which occurs on the timescale of seconds and (b) translation of this information into an actual weight change, which Precise timing of event sequences 19 occurs on the timescale of minutes or hours. The initial and immediate signal shaped by the firing rates of pre- and postsynaptic neural populations was modeled by intermediate variables we refer to as proto-weights (Gavornik et al, 2009) (Fig. 8A). Changes to the actual synaptic weights occur on a much longer timescale, and slowly converged to values determined by the proto-weights (Fig. 8B). Introducing proto-weights leads to repeated reen- forcement of learned activity patterns making them robust to spontaneous network activations. The model takes the form = −uj + ϕ (Ij(t) + Ij,syn − θ) , = 1 − pj + (pmax − 1)uj, (cid:32) N(cid:88) = −v + ϕ (cid:33) Zkuk − θv , k=1 = −γdwjkuk(t − D)(M − uj(t)) + γp(wmax − wjk)uk(t − D)uj(t), j (cid:54)= k, = wjk(t − Dp) − Wjk(t), (16) τ τf duj dt dpj dt τ dv dt τw dwjk dt τI dWjk dt where Ij,syn = Wjjuj + N(cid:88) k(cid:54)=j Wjkpkuk − Lv, τI is the time scale of the actual weights, and Dp is a time delay in the process of transforming changes in the proto-weights to changes in the actual weights. Since τw represents the timescale of changes in proto-weights, the timescale τI represents the timescale of actual synaptic weight changes. The process of translating a coincidence in firing rates into a weight change can be much longer than detecting the coin- cidence, and we can thus take τI much larger than τw (Markram and Tsodyks, 1996). It was still possible to analyze the model given in Eqs. (16), but the results were less transpar- ent. During training, the proto-weights satisfied Eq. (5). As the number of training trials increased, the proto-weights converged to the limit given in Fig. 8 A two-stage learning rule. A. Proto-weights track the signaled change to synaptic connectivity brought about by firing rate covariance. Actual weights evolve on a slower time scale and hence do not track changes in the proto-weights immediately. B. Actual synaptic weights approach proto-weight values on a longer timescale, so both eventually converge to the same value. Eq. (6). The actual weights also slowly approached the same limit; that is, Wjk converged to the limit given in Eq. (6). During replay, the proto-weights evolved according to Eqs. 16, but due to the delay Dp and the time scale τI , the actual weights re- mained unchanged and replay occurred accurately as in the Section 2.6. Modeling long term plasticity as a two-stage process allowed us to incorporate more realistic de- tails: We were able to assume that synaptic con- nections are plastic during replay, as well as during training. Replaying the sequence of neural popula- tion activations evokes long term plasticity signals through the proto-weights, and alterations to ac- tual synaptic weights do not take place until af- ter the epochs of neural activity. During reacti- vation, the actual weights already equal the val- ues necessary to elicit the timed event sequence. Then, the weakening of the connection due to LTD will precisely equal the strengthening due to LTP, resulting in no net change in the actual connec- tion weight. As long as the time scale of synaptic weight consolidation is much larger than the se- quence timescale, long term plasticity during re- firingratesynapticstrength 01proto-weight05time (s)actual weightproto-weightactual weight30time (min)synapticstrength AB 20 Alan Veliz-Cuba et al. play reinforces the learned network of weights that is already present. 5 Discussion Sequences of sensory and motor events can be en- coded in the architecture of neuronal networks. These sequences can then be replayed with cor- rect order and timing when the first element of the sequence is presented, even in the absence of any other sensory input. Experimental evidence shows that after repeated presentations of a cued training sequence, the presentation of the cue alone trig- gers a temporal pattern of activity similar to that evoked by the training stimulus (Eagleman and Dragoi, 2012; Xu et al, 2012; Shuler and Bear, 2006; Gavornik and Bear, 2014). Our goal here was to provide a biologically plausible mechanism that could govern the learning of precisely timed sequences. 5.1 Learning both the precise timing and order of sequences We demonstrated how a complex learning task can be accomplished by combining two simple mech- anisms. First, the timing of a single event can be represented by a slowly accumulating positive feed- back process (Buonomano, 2000; Durstewitz, 2003; Reutimann et al, 2004; Shea-Brown et al, 2006; Karmarkar and Buonomano, 2007; Gavornik et al, 2009; Simen et al, 2011). Second, rate dependent long term plasticity can reshape synaptic weights so that the order and precise timing of events in a sequence is encoded by the network's architec- ture (Amari, 1972; Abbott and Blum, 1996). To make the problem analytically tractable, we considered an idealized model of neural population firing rates, long term plasticity, and short term fa- cilitation. This allowed us to obtain clear relation- ships between parameters of the time-tracking pro- cess (short term facilitation) and the learning pro- cess (long term plasticity). The assumptions about model structure and parameters that were essential for sequence learning could be explicitly described in this model. Similar conditions were required for learning in more realistic models, which incorpo- rated the long timescale of LTP/LTD. A novel feature of our network model is that long term plasticity influences the length of time a neural population is active. Typical computational models of sequence learning employ networks of neurons (Jun and Jin, 2007; Fiete et al, 2010; Brea et al, 2013) or populations (Abbott and Blum, 1996) that are each active for equal amounts of time dur- ing replay. However, sensory and motor processes can be governed by networks whose neurons have a fixed stimulus tuning (Xu et al, 2012; Gavornik and Bear, 2014). Therefore, a sequence of events of varying time lengths should be represented by neu- ral populations that are each active for precisely the length of time of the corresponding event. Our model demonstrates that this can be achieved us- ing rate-based long term plasticity. 5.2 Experimental predictions The general mechanisms we described here imply a number of experimentally testable features of the neural substrates of the learning and recall of event sequences. Our analysis of the impact of noise on time encoding demonstrates a relationship be- tween the dynamical mechanism for encoding and error statistics. If time interval estimation is ac- complished through a slow process that saturates toward a threshold, then the relative error of an interval in the sequence should increase with inter- val length, and average interval estimate should be shortened during replay (Fig. 5J). This provides an innate mechanistic explanation for underestimate of time duration, contributing to existing literature that has found environmental conditions that can lead to such systematic errors (Morrone et al, 2005; Terao et al, 2008). When time is marked by a slow process that scales linearly in time, average dura- tion estimates will be close to the true estimate and relative error will be invariant to trained du- rations (Supplementary Material, Fig. 3). We sug- gest a way in which average interval estimates may Precise timing of event sequences 21 be shorter than the trained interval, if the trained interval is longer than the timescale of the slow process encoding it. If the slow process that reads out the stored time grows linearly or exponentially, average interval estimates may be nearly equal or longer than the true duration (Supplementary Ma- terial, Fig. 3). The mechanics of sequence learning could be understood further by examining the development of sequence replay accuracy with the number of trainings. Errors in sequence recall will tend to be greatest after very few trainings (Fig 4A). Corti- cal recordings reveal that, indeed, the correlation between replayed activity and training sequence evoked activity increases with the number of se- quence exposures (Eagleman and Dragoi, 2012). Additionally, our model suggests that errors in the replayed time of each event's beginning will build serially, if they rely upon a sequence of popula- tion activations. This means that errors in the total run time of a sequence will increase with sequence duration, as in Hass et al (2008). However, errors in the individual estimates of each event duration will not depend on their placement in the sequence. Such errors should decrease at a similar rate across all individual events, as suggested by our analysis in Materials and Methods. This prediction could be tested experimentally by examining how sub- jects' individual event duration estimates depend on the event's position in a learned sequence. Lastly, we predict that event sequences can be learned through rate correlation based synaptic plas- ticity acting on connection between stimulus-tuned populations. This mechanism could be probed ex- perimentally in a number of ways. First, if neu- ral activity underlying sequence learning were be- ing recorded electrophysiologically (Xu et al, 2012; Gavornik and Bear, 2014), subsequent experiments could be performed to see if electrically stimulating neurons out of sequence could disrupt learned se- quence memory. This would provide evidence that plasticity mechanisms that result from neural ac- tivity are involved in the consolidation of sequence memory. Inactivating populations in the sequence could also disrupt the replay of the remainder of the sequence, supporting our network chain model of sequence learning. For example, optogenetic meth- ods could be used to inactivate a large fraction of cells that respond to one of the events in sequence. If such a disruption were to terminate sequence re- play, this would be strong evidence for the events being represented by a chain of active populations. Furthermore, long term plasticity processes could be disrupted through the local injection of transla- tional inhibitors (Alberini, 2009). If this leads to a reduction of sequence memory robustness, it would constitute strong evidence for the importance of long term plasticity in the local circuit for sequence memory formation. 5.3 Comparison to previous models of interval timing Several previous theoretical studies have proposed neural mechanisms for the learning and recall of timed events. Models capable of representing serial event order have utilized individual units that are oscillators (Brown et al, 2000) or bistable popula- tions (Grossberg and Merrill, 1992). Recent studies have found that continuous temporal trajectories can be learned in networks of chaotic elements by training weights to downstream neurons that con- stitute a linear readout (Buonomano and Maass, 2009; Hennequin et al, 2014). A complementary approach has been used to infer time by fitting a maximum likelihood model to the rates and phases of spiking neurons in hippocampal networks (It- skov et al, 2011). Our approach is most similar to previous studies that utilize discrete popula- tions or neurons to represent serial order (Gross- berg and Merrill, 1992; Abbott and Blum, 1996; Fiete et al, 2010; Brea et al, 2013). Namely, we assume that the memory of each individual event duration is learned in parallel with the others as in Fiete et al (2010), in contrast to the serial build- ing of chains demonstrated in the model of Jun and Jin (2007). Reset models of sequence replay are supported by comparisons of human behavioral data to models that mark event durations using a clock that is reset after each event (McAuley and 22 Alan Veliz-Cuba et al. Jones, 2003), suggesting errors are made locally in time, rather than accumulated event-to-event. We have extended previous work by developing a mechanism for altering the activation time of each unit in the sequence. This learning process is dis- tinct from the approach outlined in Buonomano and Maass (2009); Hennequin et al (2014), since it solely trains the recurrent architecture between populations encoding time; tuning of a downstream readout is unnecessary. 5.4 Internal tuning of long term plasticity parameters There is a large set of parameters for which the network can be trained to accurately replay train- ing sequences. While some parameter tuning is re- quired, in simple cases we could find these param- eters explicitly. In all cases appropriate parame- ters could be obtained computationally using gra- dient descent. In biological systems plasticity pro- cesses could be shaped across many generations by evolution, or within an organism during develop- ment. Indeed, recent experimental evidence sug- gests that networks are capable of internally tuning long term plasticity responses through metaplastic- ity (Abraham, 2008). For instance, NMDA receptor expression can attenuate LTP (Huang et al, 1992; Philpot et al, 2001), while metabotropic glutamate receptor activation can prime a network for future LTP (Oh et al, 2006). We note that such mech- anisms would affect the timescale and features of LTP/LTD, not the synaptic weights themselves. 5.5 Models that utilize ramping processes with different timescales Our proposed mechanism relies on a ramping pro- cess that evolves on the same timescale as the train- ing sequence. Short term facilitation (Markram et al, 1998) as well as rate adaptation (Benda and Herz, 2003) can fulfill this role. However, other ramp- ing processes that occur at the cellular or network level are also capable of marking time (Supplemen- tary Material, Fig. 3). For instance, slow synap- tic receptor types such as NMDA can slowly in- tegrate sensory input (Wang, 2002), resulting in population firing rate ramping similar to experi- mental observations in interval timing tasks (Xu et al, 2014). Were we to incorporate slow recur- rent excitatory synapses in this way, the duration of represented events would be determined by the decay timescale of NMDA synapses. Alternatively, we could have also employed short term depres- sion as the slow process in our model. Mutual in- hibitory networks with short term depression can represent dominance time durations that depend on the network's inputs, characteristic of percep- tual rivalry statistics (Laing and Chow, 2002). This relationship between population inputs and popu- lation activity durations could be leveraged to rep- resent event times in sequences. Events that occur on much shorter or longer timescales than those we explored here could be marked by processes matched to those timescales. For instance, fast events may be represented sim- ply using synaptic receptors with rapid kinetics, such as AMPA receptors (Clements, 1996). AMPA receptor states evolve on the scale tens of millisec- onds, which would allow representation of several fast successive events. However, we would expect a lower bound on the duration of an event repre- sented by this mechanism, given by the neuronal membrane time constant (Dayan and Abbott, 2001). Slow events could also be represented by a long chain of sub-populations, each of which is activated for a shorter amount of time than the event. In the context of our model, this would mean each pop- ulation would contain sub-populations connected as a feedforward chain (Goldman, 2009). Networks of cortical neurons can have different subpopula- tions with distinct sets of timescales, due to the variety of ion channel and synaptic receptor kinet- ics (Ulanovsky et al, 2004; Bernacchia et al, 2011; Pozzorini et al, 2013; Costa et al, 2013). This reser- voir of timescales could be utilized to learn events whose timings span several orders of magnitude. Precise timing of event sequences 23 5.6 Learning the repeated appearance of an event We only considered training sequences in which no event appeared more than once (e.g. 1-2-3-4). If events appear multiple times (e.g. 1-2-1-4), then a learned synaptic weight (e.g., w21) would be weak- ened when the repeated event appears again. This can be resolved by representing each event repe- tition by the activation of a different subpopula- tion of cells. There is evidence that this occurs in hippocampal networks responding to spatial nav- igation sequences on a figure eight track (Griffin et al, 2007). Even for networks where each stimu- lus activates a specific population, sequences with repeated stimuli could be encoded in a deeper layer of the underlying sensory or motor system. The same idea can be used to create networks that can store several different event sequences contain- ing the same events (e.g. 1-2-3-4; 2-4-3-1; 4-3-2-1). If multiple sequences begin with the same event (e.g. 1-2-3-4; 1-3-2-4), evoking the correct sequence would require partial stimulation of the sequence (e.g. 1-2 or 1-3). Networks would then be less likely to misinterpret one learned sequence for another sequence with overlapping events (Abbott and Blum, 1996). 5.7 Feedback correction in learned sequences We emphasize that we did not incorporate any mechanisms for correcting errors in timing during replay. However, this could easily be implemented by considering feedback control via a stimulus that activates the population that is supposed to be active, if any slippage in event timing begins to occur. This assumes there is some external sig- nal indicating how accurately the sequence is be- ing replayed. For instance, human performance of a piece of music relies on auditory feedback sig- nals that are used by the cerebellum to correct motor errors (Zatorre et al, 2007; Kraus and Chan- drasekaran, 2010). If feedback is absent or is ma- nipulated, performance deteriorates (Finney and Palmer, 2003; Pfordresher, 2003). Similar princi- ples seem to hold in the replay of visual sequences. Gavornik and Bear (2014) showed that portions of learned sequence are replayed more accurately when preceded by the correct initial portion of the learned sequence. We could incorporate feedback into our model by providing external input to the network at several points in time, not just the ini- tial cue stimulus. 5.8 Conclusions Overall, our results suggest that a precisely timed sequence of events can be learned by a network with long term synaptic plasticity. Sequence play- back can be accomplished by a ramping process whose timescale is similar to the event timescales. Trial-to-trial variability in training and neural ac- tivity will be inherited by the sequence representa- tion in a way that depends on the learning process and the playback process. Therefore, errors in se- quence representation provide a window into the neural processes that represent them. Future ex- perimental studies of sequence recall that statis- tically characterize these errors will help to shed light on the neural mechanisms responsible for se- quence learning. Acknowledgements We thank Jeffrey Gavornik helpful comments. Funding was provided by NSF-DMS-1311755 (Z.P.K.); NSF/NIGMS-R01GM104974 (A.V-C. and K.J.); and NSF-DMS-1122094 (A.V-C. and K.J.). References Abbott LF, Blum KI (1996) Functional signif- icance of long-term potentiation for sequence learning and prediction. Cereb Cortex 6(3):406 -- 416 Abraham WC (2008) Metaplasticity: tuning synapses and networks for plasticity. Nat Rev Neurosci 9(5):387 Alberini C (2009) Transcription factors in long- term memory and synaptic plasticity. Physiol Rev 89(1):121 -- 145 24 Alan Veliz-Cuba et al. Amari SI (1972) Learning patterns and pattern se- quences by self-organizing nets of threshold ele- ments. IEEE Trans Comput 21(11):1197 -- 1206 Benda J, Herz AV (2003) A universal model for spike-frequency adaptation. Neural Comput 15(11):2523 -- 2564 Bernacchia A, Seo H, Lee D, Wang XJ (2011) A reservoir of time constants for memory traces in cortical neurons. Nat Neurosci 14(3):366 -- 72 Bienenstock EL, Cooper LN, Munro PW (1982) Theory for the development of neuron selectiv- ity: orientation specificity and binocular interac- tion in visual cortex. J Neurosci 2(1):32 -- 48 Bliss TVP, Lømo T (1973) Long-lasting potentia- tion of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. J Physiol 232(2):331 -- 356 Brea J, Senn W, Pfister JP (2013) Matching recall and storage in sequence learning with spiking neural networks. J Neurosci 33(23):9565 -- 9575 Brown GD, Preece T, Hulme C (2000) Oscillator- based memory for serial order. Psychological re- view 107(1):127 Bueti D, Buonomano DV (2014) Temporal per- learning. Timing & Time Perception ceptual 2(3):261 -- 289 Buhusi CV, Meck WH (2005) What makes us tick? functional and neural mechanisms of interval timing. Nat Rev Neurosci 6(10):755 -- 65 Buonomano DV (2000) Decoding temporal infor- mation: A model based on short-term synaptic plasticity. J Neurosci 20(3):1129 -- 1141 Buonomano DV, Maass W (2009) State-dependent computations: spatiotemporal processing in cor- tical networks. Nature Reviews Neuroscience 10(2):113 -- 125 Burgess N, Hitch GJ (1999) Memory for serial or- der: A network model of the phonological loop and its timing. Psychological Review 106(3):551 Butts DA, Weng C, Jin J, Yeh CI, Lesica NA, Alonso JM, Stanley GB (2007) Temporal pre- cision in the neural code and the timescales of natural vision. Nature 449(7158):92 -- 95 Clements J (1996) Transmitter timecourse in the synaptic cleft: its role in central synaptic func- tion. Trends Neurosci 19(5):163 -- 171 Clopath C, Busing L, Vasilaki E, Gerstner W (2010) Connectivity reflects coding: a model of voltage-based stdp with homeostasis. Nat Neu- rosci 13(3):344 -- 352 Conway CM, Christiansen MH (2001) Sequential learning in non-human primates. Trends Cogn Sci 5(12):539 -- 546 Costa RP, Sjostrom PJ, Van Rossum MC (2013) Probabilistic inference of short-term synaptic plasticity in neocortical microcircuits. Frontiers in computational neuroscience 7 Dayan P, Abbott LF (2001) Theoretical neuro- science. Cambridge, MA: MIT Press Dudek SM, Bear MF (1992) Homosynaptic long- term depression in area ca1 of hippocampus and effects of n-methyl-d-aspartate receptor block- ade. Proc Natl Acad Sci USA 89(10):4363 -- 4367 Durstewitz D (2003) Self-organizing neural inte- grator predicts interval times through climbing activity. J Neurosci 23(12):5342 -- 5353 Eagleman SL, Dragoi V (2012) Image sequence re- activation in awake v4 networks. Proc Natl Acad Sci USA 109(47):19,450 -- 19,455 Faisal AA, Selen LP, Wolpert DM (2008) Noise in the nervous system. Nat Rev Neurosci 9(4):292 -- 303 Fiete IR, Senn W, Wang CZ, Hahnloser RH (2010) Spike-time-dependent plasticity and heterosy- naptic competition organize networks to produce long scale-free sequences of neural activity. Neu- ron 65:563 -- 576 Finney S, Palmer C (2003) Auditory feedback and memory for music performance: Sound evidence for an encoding effect. Mem Cognition 31(1):51 -- 64 Fourcaud-Trocm´e N, Hansel D, Van Vreeswijk C, Brunel N (2003) How spike generation mecha- nisms determine the neuronal response to fluc- tuating inputs. J Neurosci 23(37):11,628 -- 11,640 Gardiner CW (2004) Handbook of Stochastic Methods. Springer Gavornik JP, Bear MF (2014) Learned spatiotem- poral sequence recognition and prediction in pri- mary visual cortex. Nat Neurosci 17(5):732 -- 737 Precise timing of event sequences 25 Gavornik JP, Shuler MGH, Loewenstein Y, Bear MF, Shouval HZ (2009) Learning reward timing in cortex through reward dependent expression of synaptic plasticity. Proc Natl Acad Sci USA 106(16):6826 -- 6831 Gerstner W, Kistler WM (2002) Mathematical for- mulations of hebbian learning. Biol Cybern 87(5- 6):404 -- 15 Gjorgjieva J, Clopath C, Audet J, Pfister JP (2011) A triplet spike-timing -- dependent plasticity model generalizes the Bienenstock -- Cooper -- Munro rule to higher-order spatiotem- poral correlations. Proc Natl Acad Sci USA 108(48):19,383 -- 19,388 Goldman MS (2009) Memory without feedback in a neural network. Neuron 61(4):621 -- 634 Graupner M, Brunel N (2012) Calcium-based plas- ticity model explains sensitivity of synaptic changes to spike pattern, rate, and dendritic lo- cation. Proc Natl Acad Sci USA 109(10):3991 -- 3996 Griffin AL, Eichenbaum H, Hasselmo ME (2007) Spatial representations of hippocampal ca1 neu- rons are modulated by behavioral context in a hippocampus-dependent memory task. J Neu- rosci 27(9):2416 -- 2423 Grossberg S, Merrill JW (1992) A neural network model of adaptively timed reinforcement learn- ing and hippocampal dynamics. Cognitive brain research 1(1):3 -- 38 Gutig R, Aharonov R, Rotter S, Sompolinsky H (2003) Learning input correlations through non- linear temporally asymmetric hebbian plasticity. The Journal of neuroscience 23(9):3697 -- 3714 Haider B, Hausser M, Carandini M (2013) Inhibi- tion dominates sensory responses in the awake cortex. Nature 493(7430):97 -- 100 Hass J, Blaschke S, Rammsayer T, Herrmann JM (2008) A neurocomputational model for optimal temporal processing. Journal of computational neuroscience 25(3):449 -- 464 Hausser M, Roth A (1997) Estimating the time course of the excitatory synaptic conductance in neocortical pyramidal cells using a novel volt- age jump method. The Journal of neuroscience 17(20):7606 -- 7625 Hennequin G, Vogels TP, Gerstner W (2014) Op- timal control of transient dynamics in balanced networks supports generation of complex move- ments. Neuron 82(6):1394 -- 1406 Hikosaka O, Nakamura K, Sakai K, Nakahara H (2002) Central mechanisms of motor skill learn- ing. Current opinion in neurobiology 12(2):217 -- 222 Huang YY, Colino A, Selig DK, Malenka RC (1992) The influence of prior synaptic activity on the induction of long-term potentiation. Sci- ence 255(5045):730 -- 3 Itskov V, Curto C, Pastalkova E, Buzs´aki G (2011) Cell assembly sequences arising from spike threshold adaptation keep track of time in the hippocampus. The Journal of Neuroscience 31(8):2828 -- 2834 Ivry RB, Schlerf JE (2008) Dedicated and intrin- sic models of time perception. Trends Cogn Sci 12(7):273 -- 280 Jenkins I, Brooks D, Nixon P, Frackowiak R, Pass- ingham R (1994) Motor sequence learning: a study with positron emission tomography. The Journal of Neuroscience 14(6):3775 -- 3790 Jun JK, Jin DZ (2007) Development of neu- ral circuitry for precise temporal sequences through spontaneous activity, axon remodeling, and synaptic plasticity. PLoS ONE 2(8):e723 Kandel ER (2001) The molecular biology of mem- ory storage: a dialogue between genes and synapses. Science 294(5544):1030 -- 1038 Karmarkar UR, Buonomano DV (2007) Timing in the absence of clocks: encoding time in neural network states. Neuron 53(3):427 -- 438 Kempter R, Gerstner W, van Hemmen JL (1999) Hebbian learning and spiking neurons. Phys Rev E 59:4498 -- 4514 Kleinfeld D (1986) Sequential state generation by model neural networks. Proc Natl Acad Sci USA 83(24):9469 -- 9473 Ko H, Hofer SB, Pichler B, Buchanan KA, Sjostrom PJ, Mrsic-Flogel TD (2011) Functional specificity of local synaptic connections in neo- cortical networks. Nature 473(7345):87 -- 91 26 Alan Veliz-Cuba et al. Kok P, Jehee JF, de Lange FP (2012) Less is more: expectation sharpens representations in the pri- mary visual cortex. Neuron 75(2):265 -- 270 Kraus N, Chandrasekaran B (2010) Music training for the development of auditory skills. Nat Rev Neurosci 11(8):599 -- 605 Laing CR, Chow CC (2002) A spiking neuron model for binocular rivalry. J Comput Neurosci 12(1):39 -- 53 Litwin-Kumar A, Doiron B (2012) Slow dynam- ics and high variability in balanced cortical net- works with clustered connections. Nat Neurosci 15(11):1498 -- 505 Lundstrom B (2015) Modeling multiple time scale firing rate adaptation in a neural network of local field potentials. J Comput Neurosci 38(1):189 -- 202 Ma WJ, Beck JM, Latham PE, Pouget A (2006) Bayesian inference with probabilistic population codes. Nat Neurosci 9(11):1432 -- 1438 Major G, Tank D (2004) Persistent neural activity: prevalence and mechanisms. Curr Opin Neuro- biol 14(6):675 -- 84 Malenka RC, Bear MF (2004) Ltp and ltd: an em- barrassment of riches. Neuron 44(1):5 -- 21 von der Malsburg C (1973) Self-organization of ori- entation sensitive cells in the striate cortex. Ky- bernetik 14(2):85 -- 100 Markram H, Tsodyks M (1996) Redistribution of synaptic efficacy between neocortical pyramidal neurons. Nature 382(6594):807 -- 810 Markram H, Wang Y, Tsodyks M (1998) Differ- ential signaling via the same axon of neocorti- cal pyramidal neurons. Proc Natl Acad Sci USA 95(9):5323 -- 5328 McAuley JD, Jones MR (2003) Modeling effects of rhythmic context on perceived duration: a com- parison of interval and entrainment approaches to short-interval timing. Journal of Experimen- tal Psychology: Human Perception and Perfor- mance 29(6):1102 Meyer T, Olson CR (2011) Statistical learning of visual transitions in monkey inferotemporal cor- tex. Proceedings of the National Academy of Sci- ences 108(48):19,401 -- 19,406 Miller KD (1994) A model for the development of simple cell receptive fields and the ordered arrangement of orientation columns through activity-dependent competition between on-and off-center inputs. J Neurosci 14:409 -- 441 Morrone MC, Ross J, Burr D (2005) Saccadic eye movements cause compression of time as well as space. Nat Neurosci 8(7):950 -- 4 Nabavi S, Fox R, Proulx CD, Lin JY, Tsien RY, Malinow R (2014) Engineering a memory with ltd and ltp. Nature 511(7509):348 -- 52 Oh MC, Derkach VA, Guire ES, Soderling TR (2006) Extrasynaptic membrane trafficking reg- ulated by glur1 serine 845 phosphorylation primes ampa receptors for long-term potentia- tion. J Biol Chem 281(2):752 -- 8 Oja E (1982) Simplified neuron model as a princi- pal component analyzer. J Math Biol 15(3):267 -- 273 Perin R, Berger TK, Markram H (2011) A synaptic organizing principle for cortical neuronal groups. Proc Natl Acad Sci USA 108(13):5419 -- 5424 Pfister JP, Gerstner W (2006) Triplets of spikes in a model of spike timing-dependent plasticity. J Neurosci 26(38):9673 -- 9682 Pfordresher PQ (2003) Auditory feedback in mu- sic performance: Evidence for a dissociation of sequencing and timing. Journal of Experimen- tal Psychology: Human Perception and Perfor- mance 29(5):949 Philpot BD, Sekhar AK, Shouval HZ, Bear MF (2001) Visual experience and deprivation bidi- rectionally modify the composition and func- tion of NMDA receptors in visual cortex. Neuron 29(1):157 -- 169 Pozzorini C, Naud R, Mensi S, Gerstner W (2013) Temporal whitening by power-law adap- tation in neocortical neurons. Nature neuro- science 16(7):942 -- 948 Rao RP, Sejnowski TJ (2001) Spike-timing- dependent hebbian plasticity as temporal differ- ence learning. Neural Comput 13(10):2221 -- 2237 Reutimann J, Yakovlev V, Fusi S, Senn W (2004) Climbing neuronal activity as an event- based cortical representation of time. J Neurosci Precise timing of event sequences 27 ralistic viewing. Neuron 55(3):465 -- 78 Wang XJ (2002) Probabilistic decision making by slow reverberation in cortical circuits. Neuron 36(5):955 -- 68 Xu M, Zhang SY, Dan Y, Poo Mm (2014) Repre- sentation of interval timing by temporally scal- able firing patterns in rat prefrontal cortex. Proc Natl Acad Sci USA 111(1):480 -- 485 Xu S, Jiang W, Poo Mm, Dan Y (2012) Activity recall in a visual cortical ensemble. Nat Neurosci 15:449 -- 455 Zatorre RJ, Chen JL, Penhune VB (2007) When the brain plays music: auditory -- motor interac- tions in music perception and production. Nat Rev Neurosci 8(7):547 -- 558 24(13):3295 -- 3303 Sakai K, Hikosaka O, Miyauchi S, Takino R, Sasaki Y, Putz B (1998) Transition of brain activation from frontal to parietal areas in visuomotor se- quence learning. The Journal of Neuroscience 18(5):1827 -- 1840 Shea-Brown E, Rinzel J, Rakitin BC, Malapani C (2006) A firing rate model of parkinsonian deficits in interval timing. Brain Res 1070(1):189 -- 201 Shuler MG, Bear MF (2006) Reward timing in the primary visual cortex. Science 311(5767):1606 -- 1609 Simen P, Balci F, Cohen JD, Holmes P, et al (2011) A model of interval timing by neural integration. J Neurosci 31(25):9238 -- 9253 Sjostrom PJ, Turrigiano GG, Nelson SB (2001) timing, and cooperativity jointly de- synaptic plasticity. Neuron Rate, termine cortical 32(6):1149 -- 1164 Song S, Sjostrom PJ, Reigl M, Nelson S, Chklovskii DB (2005) Highly nonrandom features of synap- tic connectivity in local cortical circuits. PLoS Biol 3(3):e68 Takeuchi T, Duszkiewicz AJ, Morris RG (2014) The synaptic plasticity and memory hypothesis: encoding, storage and persistence. Philos T Roy Soc Lond B 369(1633):20130,288 Terao M, Watanabe J, Yagi A, Nishida S (2008) Reduction of stimulus visibility compresses ap- parent time intervals. Nat Neurosci 11(5):541 -- 2 Tsodyks M, Pawelzik K, Markram H (1998) Neural networks with dynamic synapses. Neural Com- put 10(4):821 -- 835 Ulanovsky N, Las L, Farkas D, Nelken I (2004) Multiple time scales of adaptation in auditory cortex neurons. The Journal of Neuroscience 24(46):10,440 -- 10,453 Wang D, Arbib M (1990) Complex temporal se- quence learning based on short-term memory. P IEEE 78(9):1536 -- 1543 Wang X, Wei Y, Vaingankar V, Wang Q, Koepsell K, Sommer FT, Hirsch JA (2007) Feedforward excitation and inhibition evoke dual modes of firing in the cat's visual thalamus during natu-
1507.00235
1
1507
2015-07-01T13:52:07
Energy-efficient neuromorphic classifiers
[ "q-bio.NC", "cs.NE" ]
Neuromorphic engineering combines the architectural and computational principles of systems neuroscience with semiconductor electronics, with the aim of building efficient and compact devices that mimic the synaptic and neural machinery of the brain. Neuromorphic engineering promises extremely low energy consumptions, comparable to those of the nervous system. However, until now the neuromorphic approach has been restricted to relatively simple circuits and specialized functions, rendering elusive a direct comparison of their energy consumption to that used by conventional von Neumann digital machines solving real-world tasks. Here we show that a recent technology developed by IBM can be leveraged to realize neuromorphic circuits that operate as classifiers of complex real-world stimuli. These circuits emulate enough neurons to compete with state-of-the-art classifiers. We also show that the energy consumption of the IBM chip is typically 2 or more orders of magnitude lower than that of conventional digital machines when implementing classifiers with comparable performance. Moreover, the spike-based dynamics display a trade-off between integration time and accuracy, which naturally translates into algorithms that can be flexibly deployed for either fast and approximate classifications, or more accurate classifications at the mere expense of longer running times and higher energy costs. This work finally proves that the neuromorphic approach can be efficiently used in real-world applications and it has significant advantages over conventional digital devices when energy consumption is considered.
q-bio.NC
q-bio
Energy-efficient neuromorphic classifiers Daniel Mart´ı ∗ †, Mattia Rigotti ‡ † Mingoo Seok § and Stefano Fusi † ∗ D´epartement d’´Etudes Cognitives, ´Ecole Normale Sup´erieure - PSL Research University, Paris, France. Institut nationale de la sant´e et de la recherche m´edicale, France.,‡Physical Sciences Department, IBM T. J. Watson Research Center, Yorktown Heights, NY 10598,†Center for Theoretical Neuroscience, Columbia University, New York, USA, and §Department of Electrical Engineering, Columbia University, New York, USA Neuromorphic engineering combines the architectural and computational principles of systems neuroscience with semiconductor electronics, with the aim of building efficient and compact devices that mimic the synaptic and neural machinery of the brain. Neuromorphic engineering promises ex- tremely low energy consumptions, comparable to those of the nervous sys- tem. However, until now the neuromorphic approach has been restricted to relatively simple circuits and specialized functions, rendering elusive a direct comparison of their energy consumption to that used by conventional von Neumann digital machines solving real-world tasks. Here we show that a re- cent technology developed by IBM can be leveraged to realize neuromorphic circuits that operate as classifiers of complex real-world stimuli. These cir- cuits emulate enough neurons to compete with state-of-the-art classifiers. We also show that the energy consumption of the IBM chip is typically 2 or more orders of magnitude lower than that of conventional digital machines when implementing classifiers with comparable performance. Moreover, the spike-based dynamics display a trade-off between integration time and accuracy, which naturally translates into algorithms that can be flexibly deployed for either fast and approximate classifications, or more accurate classifications at the mere expense of longer running times and higher en- ergy costs. This work finally proves that the neuromorphic approach can be efficiently used in real-world applications and it has significant advantages over conventional digital devices when energy consumption is considered. neuromorphic electronic hardware VLSI technology neural networks classifica- tion Abbreviations: SVM: support vector machine — SV: support vector — RCN: ran- domly connected neuron Introduction Recent developments in digital technology and machine learning are enabling computers to perform an in- creasing number of tasks that were once solely the domain of human expertise, such as recognizing a face in a picture or driving a car in city traffic. These are impressive achieve- ments, but we should keep in mind that the human brain carries out tasks of such complexity using only a small frac- tion of the energy needed by conventional computers, the dif- ference in energy consumption being often of several orders of magnitude. This suggests that one way to reduce energy consumption is to design machines whose architecture takes inspiration from the biological brain, an approach that was proposed by Carver Mead in the late 1980s [1] and that is now known as “neuromorphic engineering”. Mead’s idea was to use very-large-scale integration (VLSI) technology to build electronic circuits that mimic the architecture of the nervous system. The first electronic devices inspired by this concept were analog circuits that exploited the subthreshold properties of transistors to emulate the biophysics of real neurons. Nowa- days the term “neuromorphic” refers to any analog, digital, or hybrid VLSI system whose design principles are inspired by those of biological neural systems [2]. Neuromorphic hardware has convincingly demonstrated its potential for energy efficiency, as proven by devices that consume as little as a few picojoules per neural event (spike) [3, 4, 5]. These devices contain however a relatively small number of elements (neurons and synapses) and they can typ- ically perform only simple and specialized tasks, making it 1–12 difficult to directly compare their energy consumption to that of conventional digital machines. The situation has changed recently with the development by IBM of the TrueNorth processor, a neuromorphic device that implements enough artificial neurons to perform com- plex real-world tasks, like large-scale pattern classification [6]. Here we show that a pattern classifier implemented on the IBM chip can achieve performances comparable to those of state-of-the-art conventional devices based on the von Neu- mann architecture. More importantly, our chip-implemented classifier uses 2 or more orders of magnitude less energy than current digital machines performing the same classification tasks. These results show for the first time the deployment of a neuromorphic device able to solve a complex task, while meeting the claims of energy efficiency contented by the neu- romorphic engineering community for the last few decades. Results We chose pattern classification as an example of a complex task because of the availability of well-established bench- marks. A classifier takes an input, like the image of a hand- written character, and assigns it to one among a set of discrete classes, like the set of digits. To train and evaluate our clas- sifiers we used three different datasets consisting of images of different complexity (see Fig. 1a). We start by describing the architecture of the classifier that we plan to implement on the neuromorphic chip. The classifier is a feed-forward neural network with three layers of neurons, and it can be simulated on a traditional digital com- puters. We will call this network the ‘neural classifier’ to dis- tinguish it from its final chip implementation, which requires adapting the architecture to the connectivity constraints im- posed by the hardware. The neural classifier also differs from the final hardware implementation in that it employs neu- rons with a continuous activation function, whereas the IBM neuromorphic chip emulates spiking neurons. Despite the dif- ferences, the functionality of the neural classifier and its final chip implementation is approximately the same, as we show below. We list the procedure for adapting the architecture of the neural classifier into its chip implementation as a contribu- tion in its own right, since it can be directly extended for the implementation of generic neural systems on other hardware substrates. Architecture of the neural classifier Figure 1b illustrates the three-layer neural classifier. The first layer encodes the pre- processed input and projects to the neurons in the interme- diate layer through connections with random weights. Each of these Randomly Connected Neurons (RCNs) receives there- fore a synaptic current given by a randomly weighted sum of the inputs, which the RCNs transform into activation levels in a non-linear way—in our case, through a linear rectifica- tion function: f (x) = x if x > 0, and 0 otherwise. The combination of a random mixing of the inputs together with a non-linear input-output transformation efficiently expands the dimensionality of the resulting signal (see e.g. [7, 8, 9]), thereby increasing the chances that downstream neurons can Figure 1. Datasets, architecture of the classifier, architecture of a single core, and chip implementation. a Samples of the three datasets used to evaluate the performance of our classifier. MNIST contains handwritten digits (10 classes); MNIST-back-image contains the digits of the MNIST dataset on a background patch extracted randomly from a set of 20 images downloaded from the Internet; LATEX contains distorted versions of 293 characters used in the LATEX document preparation system. For more details about the datasets, see Methods. b Architecture of the neural network classifier. The images to classify are preprocessed (see Methods) and represented as patterns of activity of a population of Nin input neurons (left, black dots). These input neurons send random projections to a second layer of N Randomly Connected Neurons (RCNs) (green circles), which transform nonlinearly their synaptic inputs into firing activity. The activity of the RCNs is then read out by a third layer of neurons, each of which is trained to respond to only one class (red circles). c Architecture of a single core in the chip. Horizontal lines represent inputs, provided by the axons of neurons that project to the core. Vertical lines represent the dendrites of the neurons in the core (one dendrite per neuron). Active synapses are shown as dots in a particular axon-dendrite junction. The synaptic input collected by the dendrites is integrated and transduced into spike activity at the soma (filled squares on top). The spikes emitted by the neuron are sent via its axon to a particular input line, not necessarily on the same core. Blue lines represent the flow of input and output signals. The panel includes an example of internal connection: the upmost axon carries the output activity of the leftmost neuron in the core (other connections are left unspecified). d Implementation of the neural network classifier in a chip with connectivity constraints. The input is fed into all the cores in the RCN layer (shaded blue), whose neurons project to the input lines of readout cores (shaded yellow) in a one-to-one manner (green curves). The outputs of the readout units are combined together off-line to generate the response of the output neuron (shaded red). See the main text for the description of the different modules. discriminate signals belonging to distinct classes. This dis- crimination is carried out by a set of output units in the last layer, which compute a weighted sum of the RCNs activity. The weights are trained so that each output unit responds to one separate class (one-vs-all code). Details are given in the Methods. Once the network is trained, a class is assigned to each input patterns according to which output unit exhibits the highest activation. Chip implementation of the neural classifier We implemented the neural classifier on the IBM neuromorphic chip described in [10, 6]. The first step of the conversion of the abstract neu- ral classifier to an explicit chip implementation is the transfor- mation of the input patterns into a format that is compatible with the spike-based coding of the TrueNorth system. For this we simply employ a firing rate coding and convert the integer value of every input component to a spike train with a propor- tional number of spikes, a prescription that is commonly used in neurocomputational models such as the Neural Engineering framework [11]. Specifically, input patterns are preprocessed and formatted into 256-dimensional vectors representing the firing activity of the input layer (the same preprocessing step was applied in the neural classifier, see Methods). This vector of activities is then used to generate 256 regular-firing spike trains that are fed into a set of cores with random and sparse connectivity. This set of cores constitutes the RCN layer. Like in the neural classifier, the neurons in the RCN layer receive synaptic inputs that consist of randomly weighted combina- tions of the input, and transform their synaptic inputs into firing activity according to a nonlinear function. On the chip this function is given by the neuronal current-to-rate transduc- tion, which approximates a linear-rectification function [12]. Discriminating the inputs coming from the RCN layer re- quires each output unit to read from the whole layer of RCNs, which in our implementation contains a number of neurons N that can be as large as 214. Moreover, all the readout connec- titions have to be set at the weights computed by the training procedure. These requirements exceed the constraints set by the chip design, in terms of the maximal number of both in- coming and outgoing connections per neuron, as well as the resolution and the freedom with which synaptic weights can In this paragraph we will present a set of prescrip- be set. 2 tions that will allow us to circumvent these limitations, and successfully instantiate our neural classifier on the IBM sys- tem. The prescriptions we are presenting are specific to the TrueNorth architecture, but the types of constraints that they solve are shared by any physical implementation of neural systems, whether it is biological or electronic. It is therefore instructive to discuss the constraints and the prescriptions to obviate them in detail, as they can be easily extended to other more generic settings. 1. Constraints on connectivity. The IBM chip is organized in cores, each of which contains 256 integrate-and-fire neurons and 256 input lines that intersect with one another forming a crossbar matrix of programmable synapses (Fig. 1c). Each neuron can connect to other neurons by projecting its axon (output) to a single input line, either on the same core or on a different core. With this hardware design the maximum number of incoming connections per neuron, or fan in, is 256. Likewise, the maximal number of outgoing connections per neuron, or fan out, is 256, each of which are restricted to target neurons within a single core. 2. Constraints on synaptic weight precision. Synapses can be either inactive or active. The weight of an active synapse can be selected from a set of four values given by signed integers of 9-bit precision. These values can differ from neuron to neu- ron. Which of the four values is assigned to an active synapse depends on the input line: all synapses on the same input line are assigned an index that determines which of the four val- ues is taken by each synapse (e.g. if the index assigned to the input line is 2, all synapses on the input line take the second value of the set of four available synaptic weights, which may differ from neuron to neuron). The design constraints that we just described can be over- come with the following set of architectural prescriptions. P1. Overcoming the constraints on connectivity. We intro- duced an intermediate layer of neurons, each of which inte- grates the inputs from 256 out of the total N RCNs. Accord- ingly, the firing rates of these intermediate neurons represent a 256/N portion of the total input to an output unit. These partial inputs can then be combined by a downstream neu- ron, which will have the same activity as the original output 9876543210inputoutput2randomconnectionsreadoutconnectionsnonlinearunits(RCNlayer)RCNlayerreadoutoutputinput+++adbcnaxons(inputs)nneurons(outputs)MNISTMNIST-back-imageLATEX unit. If the total number of the partial inputs is larger than the total number of incoming connections of the neurons that represent the output units (in our case 256), the procedure can be iterated by introducing additional intermediate layers. The final tree will contain a number of layers that scales only logarithmically with the total number of RCNs. For simplic- ity we did not implement this tree on chip and we summed off-chip the partial inputs represented by the firing activity of the readout neurons. Notice also that this configuration requires readout neurons to respond approximately linearly to their inputs, which can be easily achieved by tuning read- out neurons to operate in the linear regime of their current- to-rate transduction function (i.e., the regime in which their average input current is positive). This procedure strongly relies on the assumption that information is encoded in the firing rates of neurons; if the spiking inputs happen to be highly synchronized and synchronization encodes important information, this approach would not work. P2. Overcoming the constraints on synaptic weight preci- sion. Reducing the weight precision after learning usually only causes moderate drops in classification performance. For example, in the case of random uncorrelated inputs, the scal- ing properties of the capacity of the classifier (i.e., number of classes that can be correctly classified) remain unchanged, even when the number of states of the synaptic weights is reduced to two [13]. Instead, the performance drop is catas- trophically larger when the weight precision is limited also during learning [14, 15] and in some situations the learning problem becomes NP-complete [16] In our case the readout weights are determined off-chip, using digital conventional computers that operate on 64 bit numbers, and then quan- tized in the chip implementation. The performance drop is almost negligible for a sufficient number of synaptic levels. In our case we quantized the readout weights of the original classifier on an integer scale between −28 and 28. Each quan- tized weight was then implemented as the sum of four groups of 6 synaptic contacts, where each contact in the group can either be inactive (value 0) or activated at one of the 6 values: ±1,±2,±4. The multiplicity of this decomposition (19 can be for instance decomposed as (4) + (1 + 4) + (1 + 4) + (1 + 4) or (2)+(2+4)+(2+4)+(1+4)) is resolved by choosing the decom- position that is closest to a balanced assignment of the weights across the 4 groups (e.g. 19 = (4) + (1 + 4) + (1 + 4) + (1 + 4)). This strategy requires that each original synapse be repre- sented by 24 synapses. We implemented this strategy by replicating each readout neuron 24 times and by distributing each original weight across 24 different dendritic trees. These synaptic inputs are then summed together by the off-line sum- mation of all readout neuron activities that correspond to the partial inputs to a specific output unit (see Methods for de- tails). A similar strategy can be used to implement networks with synaptic weights that have even a larger number of lev- els and the number of additional synapses would scale only logarithmically with the total number of synaptic levels that is required. However, it is crucial to limit individual synapses to low values, in order to avoid synchronization between neu- rons. This is why we limited to 4 the maximum synaptic value of individual synapses of the chip. Classification performance and speed-accuracy trade-off Our neuromorphic classifier implemented on the TrueNorth chip was emulated on a simulator developed by IBM. As the TrueNorth chip is entirely digital, the simulator reproduces exactly the behavior of the chip [10]. In Fig. 2a we show the dynamics of two typical runs of the simulator classifying images from the MNIST-back-image dataset. Upon image presentation, the RCNs in the intermediate layer start inte- Figure 2. The neuromorphic classifier in action. a Spikes emit- ted by readout neurons during an easy (top) and a difficult (bottom) classification, after removing the trend caused by the intrinsic constant currents. Each curve corre- sponds to the readout output associated with the digit indicated by the color code. Samples are drawn from the MNIST-back-image dataset. b Test error as a function of classification time (i.e., the time over which spikes are integrated) and energy. The error is averaged over the first 1000 test samples of the MNIST (red) and MNIST- back-image (blue) datasets. Each dashed horizontal line indicates the best test error achieved with support vector classifiers for a given dataset, based on the evaluation of the whole test set. c Classification times for different thresholds in spike differ- ence (as indicated in the legend), for the MNIST and MNIST-back-image datasets. For each threshold we plot all classification times (thin lines) as well as the sample mean (shorter ticks on top). The performances associated with each threshold are indicated in the y-axis. When the threshold in spike difference is infinite (black), the classification is assessed at t = 500 ms (i.e., there is no stopping criterion). In all panels the chip uses N = 16384 RCNs. grating the input signal (not shown) and, a few tens of mil- liseconds later, they start emitting spikes, which are passed to the readout neurons. The figure shows the total number of spikes emitted by the readout neurons since input activation, after subtracting the overall activity trend caused by baseline activity. For simple classifications, in which the input is easily rec- ognizable, the readout neuron associated with the correct class is activated in less than 100 ms (Fig. 2a, top). More diffi- cult cases require the integration of spikes over longer time intervals, as the average synaptic inputs to different readout neurons can be very similar (Fig. 2a, bottom). This sug- gests that the performance of the classifier, as measured by the classification error rate on the test set, should improve with longer integration intervals. This trade-off between speed and performance is illustrated in Fig. 2b, which shows the classification performance versus elapsed time for the MNIST and MNIST-back-image datasets. The performance increases monotonically with time until it saturates in about half a second, with a highest performance of 97.27% for MNIST (98.2% with 10-fold bagging), and 77.30% for MNIST-back- image. These performances are not too far from the best classification results achieved so far: 99.06% for MNIST (us- ing maxout networks on the permutation invariant version of the MNIST dataset, which does not exploit any prior knowl- edge about the two-dimensional structure of the patterns [17]) and 77.39% for MNIST-back-image (with support vector clas- sifiers [18], although methods combining deep nets, feature 3 05010000.51performance(%)energyconsumption(mJ)MNISTMNIST-back-imageb0250500time(ms)69.977.987.997.20200400performance(%)classificationtime∞80402010spikedifferenceMNISTc59.667.675.078.40200400classificationtimeMNIST-back-image05000123456789−25002500500time(ms)emittedspikes,detrendeda learning, and feature selection can achieve performances as high as 87.75% [19]). Energy-speed-accuracy trade-off As just discussed, accuracy has a cost in term of energy because longer integration times entail more emitted spikes per classification and a larger base- line energy costs, which in our case is the dominant contri- bution to the total energy consumption. We estimated the energy consumption as described in section 5 and we found that the energy per classification never exceeds 1 mJ for our network configuration. With the energy needed to keep lit a 100 W light bulb for a second, one could perform 105 classi- fications, which is equivalent to around one classification per second uninterruptedly for almost one day. Notice that this estimate is based on a classification that lasts 0.5 s and, there- fore, does not take into account the fact that most patterns are correctly classified in a significantly shorter time (see Fig. 2a, top). If the integration and emission of spikes is stopped as soon as one of the output units is significantly more active than the others, then the average energy consumption can be strongly reduced. The criterion we used to decide when to stop the integration of spikes (and thus the classification) was based on the spikes emitted by the readout units. Specifically, we monitored the cumulative activity of each output unit by counting all the spikes emitted by the corresponding readout neurons. We stopped the classification when the accumulated activity of the leading unit exceeded that of the second unit by some threshold. The decision was the class associated with the leading output unit. In Fig. 2c we show the performances and the correspond- ing classification times for several thresholds. Low thresholds allow for faster yet less accurate classifications. In both the MNIST and MNIST-back-image datasets, the patterns that require long classifications times are rare. While the perfor- mance barely changes for large enough thresholds, the average classification time can be substantially reduced by lowering the threshold. For example, for the MNIST dataset the clas- sification time drops by a factor of 5 (from 500 ms to 100 ms) and, accordingly, so does the energy consumption (from 1 mJ to 0.2 mJ). Faster classifications are also possible by increas- ing either the average firing rate or the total number of RCNs, both of which entail an increase in energy consumption, which might be partially or entirely compensated by the decrease in the classification time. These expedients will speed up the integration of spike-counts and, as a result, the output class will be determined faster. In all cases both the energy cost and the classification per- formance increase with the total number of emitted spikes or, equivalently, with integration time, if the average firing rate is fixed. This is a simple form of a more general energy-speed- accuracy trade-off, a phenomenon that has been described in several biological information-processing systems (e.g. [20]), and that can confer great functional flexibility to our classi- fier. One advantage of basing the computation on a temporal accumulation of spikes is that the classifier can be interrupted at any time at the cost of reduced performance, but with- out compromising its function. This is in stark contrast to some conventional clock-based centralized architectures whose mode of computation crucially relies on the completion of en- tire monolithic sets of instructions. We can then envisage uti- lization scenarios where a spiking-based chip implementation of our classifier is required to flexibly switch between precise long-latency classifications (like, e.g., those involving the iden- tification of targets of interest) and rapid responses of limited accuracy (like the quick avoidance of imminent danger). Notice that both the simulated and implemented net- works, although entirely feed-forward, exhibit complex dy- 4 namics leading to classification times that depend on the dif- ficulty associated with the input. This is because neurons are spiking and the final decision requires some sort of accumu- lation of evidence. When a stimulus is ambiguous, the units representing the different decisions receive similar inputs and the competition becomes harder and longer. This type of be- havior is also observed in human brains [21]. We will now focus on the comparison of energy consump- tion and performance between the neuromorphic classifier and more conventional digital machines. Energy consumption and performance: comparison with con- ventional digital machines We compared both the classifica- tion performance and the energy consumption of our neuro- morphic classifier to those obtained with conventional digi- tal machines implementing Support Vector Machines (SVMs). SVMs offer a reasonable comparison because they are among the most successful and widespread techniques for solving machine-learning problems involving classification and regres- sion [22, 23, 24], and because they can be efficiently imple- mented on digital machines. To better understand how the energy consumption scales with the complexity of the classification problem, it is useful to summarize how SVMs work. After training, SVMs classify an input pattern according to its similarity to a set of tem- plates, called the support vectors, which are determined by the learning algorithm to define the boundaries between classes. The similarity is expressed in terms of the scalar product be- tween the input vectors and the support vectors. As argued above, we can improve classification performance by embed- ding the input vectors in a higher-dimensional space before classifying them. In this case SVMs evaluate similarities by computing classical scalar products in the higher-dimensional space. One of the appealing properties of SVMs is that there is no need to compute explicitly the transformation of inputs into high-dimensional representations. Indeed, one can skip this step and compute directly the scalar product between the transformed vectors and templates, provided that one knows how the distances are distorted by the transformation. This is known as the “kernel trick” because the similarities in a high- dimensional space can be computed and optimized over with a kernel function applied to the inputs. Interestingly, the ker- nel associated with the transformation induced by the RCNs of our neural classifier can be computed explicitly in the limit of a large number of RCNs [18]. This is also the kernel that we used to compare the performance of SVMs against that of our neural classifier. Unfortunately, classifying a test input by computing its similarity to all support vectors becomes unwieldy and com- putationally inefficient for large datasets, as the number of support vectors typically scales linearly with the size of the training set in many estimation problems [25]. This means that the number of operations to perform, and hence the en- ergy consumption per classification, also scales with the size of the training set. This makes SVMs and kernel methods com- putationally and energetically expensive in many large-scale tasks. In contrast, our neural network algorithm evaluates a test sample by means of the transformation carried out by the RCNs. If the RCN layer comprises N neurons and the input dimension is Nin, evaluating the output of a test sam- ple requires O(Nin · N ) synaptic events. Thus for large sam- ple sizes, evaluating a test sample in the network requires far fewer operations than when using the “kernel trick”, because the number is effectively independent of the size of training set (cfr. [26, 27]). Systems such as ours may therefore display considerable energy advantages over SVMs when datasets are large. Figure 3. Energy-accuracy trade-off. a,c Dependence of the classification accuracy on the number of Randomly Connected Neurons (RCNs) in the neural classifier and on the number of support vectors (SVs) in the SVC. Panel a shows this dependence for the MNIST dataset, and panel b for the MNIST-back-image. As the number of RCNs increases, the classifier becomes more accurate at the cost of higher energy consumptions (b,d). The energy consumption is based on the average time it takes to the neural classifier to perform the classification (see Fig. 2c). We also show the performance achieved by three different implementations of support vector classifiers (legend code: SVC, libsvm; rSVC, reduced primal; SVC; SVCperf, cutting plane subspace pursuit). The algorithms rSVC, SVCperf minimize the number of support vectors (SVs) with respect to the optimal value and reduce, therefore, the energy consumption levels at test time. The number of SVs used by the standard algorithm (libsvm), on the other hand, can go beyond the optimal value by reducing sufficiently the soft-margin parameter and pushing the classifier to overfit the data. In all cases, the energy consumption increases linearly with the number of SVs, as the number of operations per classification at test time scales linearly with the number of SVs. The vertical thin lines indicate the abscissa at best performance for the IBM chip (red) and SVM implementations (black). For reference we indicate the best performance achieved by the chip with a horizontal dashed line. The horizontal arrow indicates the reduction in energy consumption that would be attained if the efficiency of digital machines reached the theoretical lower bound estimated by [28]. The relation between number of SVs and energy consumption was determined by simulating the i7 Intel chip running a program that implements an SVM at test time. b Same as a, but on the MNIST-back-image dataset. In both cases our neuromorphic classifier exhibits an energy cost per classification that is orders of magnitude smaller. In Fig. 3 we compare the energy consumption and per- formance of the neuromorphic classifier to those of an SVM implemented on a conventional digital machine. More specif- ically, we estimated the energy expenditure of a digital SVM using a simulator of the Intel i7 processor, which was the ma- chine with the best energy performance among those that we simulated (see Methods section 5 and Discussion). The en- ergy cost per support vector per pattern was estimated to be around 5.2 µJ, a quantity that is not far above what is con- sidered as a lower bound on energy consumption for digital machines [28]. For both the neuromorphic classifier and the digital SVM we progressively increased the performance of the classifiers by increasing the number of RCNs (in the case of the neuromorphic classifier), and by varying the number of support vectors (in the case of the SVM), see Figures 4a,c. For the SVM we tried three different algorithms to minimize the number of support vectors and hence the energy consump- tion (for more details, see caption of Fig. 3 and Methods). For the IBM chip we estimated the energy consumption both in the case in which we stopped the classifications with the criterion described in the previous section and in the case in which the classification time was fixed at 500 ms (see Fig. 7 in Suppl. Info.). In both cases the energy consumption is signif- icantly lower for the neuromorphic classifier, being in the for- mer case approximately 2 orders of magnitude smaller for both the MNIST and the MNIST-back-image datasets, while still achieving comparable maximal performances (Fig. 4b,d). Scalability The MNIST dataset only has 10 output classes. We wondered whether the advantage of the neuromorphic clas- sifier in terms of energy consumption is preserved when the number of classes increases and the classification task becomes more complex. To study how the energy consumption scales with the number of classes we used the LATEX dataset, which contains 293 classes of distorted characters. We progressively increased the number of classes to be learned and classified and we studied the performance and the energy consumption of both the digital implementation of the SVM and the neuro- morphic classifier. Specifically, given a number of classes that was varied between 2 and 293, we selected a random subset of all the available classes, and we trained both the SVM and the neural classifier on the same subset. The results are averaged over 10 repetitions, each one with a different sample of output classes. To make a meaningful comparison between the the en- ergy consumed by a SVM and the neuromorphic classifier, we equalized all the classification accuracies, as follows. For each classification problem we varied the margin penalty param- eter of the standard SVC using grid search and picked the best performance achieved. We then varied the relevant pa- rameters of the other two classifiers so that their classification accuracy matched or exceeded the accuracy of the standard SVC. Specifically, we progressively increased the number of basis functions (in the primalSVC method) and the number of RCNs (in the neural classifier) until both reached the tar- get performance. For each classification problem we averaged over 10 realizations of the random projections of the neural classifier. The results are summarized in Fig. 4. The energy con- sumption is about two orders of magnitude larger for the SVM throughout the entire range of variation of the num- ber of classes that we considered, although for a very small (2–3) number of classes the advantage of the neuromorphic classifier strongly reduces, most likely because the algorithms to minimize the number of SVs work best when the number 5 1011021031046080100numberofRCNsorsupportvectorsperformance(%)networkchipaMNISTlibsvmSVCperfprimalSVC255075101102103104performance(%)numberofRCNsorsupportvectorsnetworkchipcMNIST-back-imagelibsvmSVCperfprimalSVCb60801000.010.1110100averageenergyconsumption(mJ)networkchiplibsvmSVCperfprimalSVC2550750.010.1110averageenergyconsumption(mJ)networkchipdlibsvmSVCperfprimalSVC of classes is low. This plot indicates that the energy advan- tage of the neuromorphic classifier over SVMs implemented on conventional digital machines is maintained also for more complex tasks involving a larger number of classes. It is interesting to discuss the expected scaling for grow- ing number of classes. Consider the case of generic C classes multi-class problems solved through reduction with multiple combined binary SVMs. In a one-vs-all reduction scheme, each binary classifier is trained to respond to exactly one of the C classes, and hence C SVMs are required. For each SVM, one needs to compute the scalar products between the test sample to be classified and the NSV support vectors. Each scalar product requires Nin multiplications and sums. In the favorable case in which all binary classifiers happen to share the same support vectors, the scalar products can be com- puted only once and would require Nin·NSV operations. These NSV scalar products then need to be multiplied by the cor- responding coefficients, which are different for the different SVMs. This requires additional CNSV operations. If NSV scales linearly with C, as in the cases we analyzed, then the total energy E will scale as E ∼ NinC + C 2. When C is small compared to Nin, the first term dominates, and the expected scaling is linear. However, for C > Nin the scaling is expected to be at least quadratic. It can grow more rapidly if the support vectors are different for different classifiers. Interestingly the expected scaling for the neural network classifiers that we considered is the same. The energy con- sumption mostly depends on the number of needed cores. This number will be proportional to the number of RCNs, N , mul- tiplied by the number of classes. Indeed, each core can receive up to 256 inputs, so the total number of needed cores will be proportional to (cid:100)N/256(cid:101), with (cid:100)·(cid:101) denoting the ceiling func- tion. Moreover, the number of readout units, which are the output lines of these cores, will be proportional to the num- ber of classes. Hence the N C dependence. In the cases we analyzed N depends linearly on the number of classes, and hence the energy depends quadratically on C, as in the case of the SVMs when C is large enough. Notice that the there is a second term which also scales quadratically with C that contributes to the energy. The second term comes from the necessity of replicating the RCNs C times, due to the limited fan out of the RCNs. Again, under the assumption of N ∼ C, also this term will scale quadratically with C. Given that the scaling with the number of classes is basi- cally the same for the neuromorhic classifier as for the SVMs, it is not unreasonable to hypothesize that the energy consump- tion advantage of the neuromorphic implementation would be preserved also for a much larger number of classes. Discussion Our results indicate that neuromorphic devices are mature enough to achieve performances on a real-world machine- learning task that are comparable to those of state-of-the-art conventional devices with von Neumann architecture, all just by using a tiny fraction of their energy. Our conclusions are based on a few significant tests, based on a comparison limited to our neuromorphic classifier and a few digital implementa- tions of SVMs. This clearly restricts the generality of our results and does not preclude situations in which the advan- tage of the neuromorphic approach might be less prominent. In any case, the merit of our study is to offer a solid com- parison with implementations on current conventional digital 6 platforms that are energy-efficient themselves. In particular, the algorithm we used on conventional digital machines in- volves only multiplications between matrices and vectors, the efficiency of which has been dramatically increased in the last decades thanks to optimized parallelization. Furthermore, not only we tried to match the classification performance of the competitors, but we also considered two additional SVM al- gorithms that minimize the number of support vectors, and hence the final number of operations. Other choices for SVM algorithms would certainly lead to different estimates for en- ergy consumption, but it is rather unlikely that they would change across 2 orders of magnitude. It is possible that full custom unconventional digital machines based, e.g., on field programmable gate arrays (FPGAs) would be more energy- efficient, but it is hard to imagine that they would break the predicted energy wall discussed in [28]. If this assumption is right, neuromorphic hardware would always be more efficient when performing the type of tasks that we considered. More- over, analog neuromorphic VLSI or unreliable digital tech- nologies might allow for a further reduction of energy con- sumption, probably by another order of magnitude [5, 29, 30]. The current energy consumption levels achieved by analog sys- tems are very close to those of biological brains in terms of en- ergy per spike, although many of these systems are relatively small and it is unclear whether they can ever be extended to brain-scale architectures. Other custom chips that can solve real-world tasks have been designed. An example is the FPGA chip NeuFlow, de- signed to implement convolutional networks for visual recog- nition. The chip is digital and uses as little as 4.9 × 1011 operations/W or, equivalently, 2 pJ/operation. It is also interesting to discuss the performance of other conventional digital processors in the benchmarks we exam- ined. Let us consider for example the implementations of SVMs classifying the MNIST digits with about 104 support vectors, which is roughly the number of vectors we need to achieve the best classification accuracy. As we have shown, the Intel i7 takes about 10 ms to perform a classification, at an approximate cost of 50 mJ. The IBM chip, in contrast, re- quired 1 mJ for the longest classification times (500 ms), and 0.2 mJ for the average classification time (100 ms). We also quantified the energy cost of the ARMv7, which is a more energy-efficient yet slower microprocessor often used in mo- bile technologies. Its energy consumption per classification was substantially higher, around 700 mJ. The main reason for this high consumption is that it takes more than 0.6 seconds to perform a single classification. And the baseline consumption, which increases linearly with the classification time, is a large portion of the total energy needed for a classification. Finally, we considered the recent Xeon Phi, which has a massively par- allel architecture and is employed in high performance com- puting applications. As we do not have a simulator for the Phi, we could only indirectly estimate a lower bound for the energy consumption (see Methods for more details). Accord- ing to our estimate, a single classification requires only 0.2 µs and uses about 16 mJ, which would be significantly lower than the energy cost of the i7 and very close to the estimated lower limit of energy consumption [28], but still larger than the con- sumption of the IBM chip. Notice however that both the clas- sification time and the energy consumption of the Xeon Phi processor are very likely to be grossly underestimated, as they are simply derived from the peak performance of 100 Tflop/s. The estimates for the i7 and the ARMv7 are significantly more reliable, because we derived them by simulating the proces- sors. To summarize, our results compellingly suggest that the neuromorphic approach is finally competitive in terms of en- Figure 4. Dependence of the energy consumption on the number of classes a Classification accuracy for the neural classifier and for two SVM algorithms, as a function of the number C of classes for the LATEX dataset. The parameters of the different classifiers are tuned to have approximately the same classification accuracy. b Energy consumption as a function of the number of classes, for the LATEX dataset. Given a number C of classes, every point in the plot is obtained by training a given classifier on C randomly sampled classes among the 293 available ones. This procedure is repeated 10 times for every value of C and every type of classifier. Each datapoint associated with the neural classifier (’RCN’) was in turn estimated from a sample of 10 realizations of the random connections (squares indicate sample means, errorbars indicate the 0.1 and 0.9 fractile of the sample). c As in b, but number of support vectors and RCNs as a function of the number of classes. Dataset MNIST MNIST-back-image LATEX size image 28 × 28 28 × 28 32 × 32 num. classes size trai- ning set 10 10 293 60000 12000 14650 size test set 10000 50000 9376 ergy consumption in useful real-world machine learning tasks and constitutes a promising direction for future scalable tech- nologies. The recent success of deep networks for large-scale machine learning [31, 32] makes neuromorphic approaches particularly relevant and valuable. This will be certainly true for neuromorphic systems with synaptic plasticity, which will enable these devices to learn autonomously from experience. Learning is now available only in small neuromorphic systems [33, 34, 35], but hopefully new VLSI technologies will allow us to implement it also in large-scale neural systems. ACKNOWLEDGMENTS. This work was supported by DARPA SyNAPSE, Gatsby Charitable Foundation, Swartz Foundation and Kavli Foundation. We are grateful to the IBM team led by Dr. D. Modha for their assistance with the IBM chip simulator. In particular we thank John Arthur and Paul Merolla for their help with the estimate of the chip energy consumption. DM acknowledges the support from the FP7 Marie Curie Actions of the European Commission and the ANR-10-LABX-0087 IEC and ANR-10-IDEX-0001-02 PSL grants. Materials and Methods Images sets for classification benchmarksWe used three datasets in our study: MNIST, MNIST-back-image, and LATEX. The MNIST dataset consists of images of handrwit- ten digits (10 classes) [36]. The MNIST-back-image dataset contains the same digits of MNIST, but in this case the back- ground of each pattern is a random patch extracted from a set of 20 black and white images downloaded from the In- ternet [37]. Patches with low pixel variance (i.e. containing little texture) are discarded. The LATEX dataset consists of distorted versions of 293 characters used in the LATEX docu- ment preparation system [38, 39]. All datasets consist of l × l pixel gray-scale images, and each of such pixel images is asso- ciated with one out of C possible classes. The size of the pixel images, the number of classes, and the sizes of the training and test sets depend on the data set (see table below). Preprocessing Every sample image was reshaped as a l2- dimensional vector, and the average gray level of each compo- nent was subtracted from the data. The dimensionality of the resulting image vector was then reduced to 256 using PCA. To guarantee that all the selected components contributed uni- formly to the patterns, we applied a random rotation to the principal subspace (see, e.g., [40]). We denote by Nin = 256 the dimension of that subspace. The architecture of the network and the training algorithm We map the preprocessed Nin-dimensional vector image, s, into a higher dimensional space through the transformation xi = f (wi · s), i = 1, . . . , N, where wi is an Nin-dimensional sparse random vector and f (·) is a nonlinear function. This is the transformation induced by a neural network with Nin input units and N output units with activation function f (·). More succinctly, x = f (WT s), [ 1 ] where W is a weight matrix of dimensions Nin × N formed by adjoining all the column weight vectors wi, and where f (·) acts componentwise, i.e., f (x) ≡ (f (x1), . . . , f (xNin ))T . The output of the random nonlinear transformation, x, is used as the input to a linear NC -class discriminant, consist- k=1 Jjkxk, with ing of NC linear functions of the type yj =(cid:80)N j = 1, . . . , C. More compactly, y = Jx, [ 2 ] where y = (y1, . . . , yC )T , J is a C×N matrix, and x is given by Eq. [ 1 ]. A pattern x is assigned to class Cj if yj(x) > yk(x) for all j (cid:54)= k. The elements of J are learned offline by impos- ing a 1-of-NC coding scheme on the output: if the target class is j then the target output t is a vector of length NC where all components are zero except component tj, which is 1. For the offline training of weights we use the pseudoinverse, which minimizes the mean squared error of the outputs. This tech- nique has been shown to be a good replacement for empirical minimization problems when the dataset is embedded in a ran- dom high-dimensional space, which is our case [41, 26, 42, 27]. Neuromorphic chip implementation The chip is composed of multiple identical cores, each of which consists of a neuromorphic circuit that comprises n = 256 ax- ons, n neurons, and n2 adjustable synapses ([43, 10, 6], see also Fig. 1c). Each axon provides the inputs by feeding the spiking activity of one given neuron that may or not reside in the core. The incoming spiking activity to all n axons in a core is represented by a vector of activity bits (A1(t), . . . , An(t)) whose elements indicate whether or not the neurons associ- ated with the incoming axons emitted a spike in the previ- ous time step. The intersection of the the n axons with the n neurons forms a matrix of programmable synapses. The weight of active synapses is determined by the type of axon and the type of neuron the synapse lies on. Specifically, each core can contain up to four different types of axon, labeled Gj = {1, 2, 3, 4}, whereas it can accommodate an unlimited 7 1010000.51numberofclassesperformanceRCNlibsvmprimalSVC101000.991a0.010.111010100energyconsumption(mJ)numberofclassesRCNprimalSVClibsvmb10100101102103104numberofclassesnumberSVs,RCNsRCNlibsvmprimalSVCc i, . . . , S4 i number of neuron types, each of which having four associated synaptic weights Si = (S1 i ). The strength of an active synapse connecting axon j with neuron i is SGj , that is, the axon type determines which weight to pick among the weights associated with neuron i. The net input received by neuron i i BijAj(t), where Bij is 1 or 0 depending on whether the synapse between axon j and dendrite i is active or inactive. at time step t is therefore hi(t) =(cid:80)n j=1 SGj At each time step the membrane potential Vi(t) of neu- ron i receiving input h(t) is updated according to Vi(t + 1) = Vi(t) − β + hi(t), where β is a constant leak. If Vi(t) becomes negative after an update, it is clipped to 0. Conversely, when Vi(t) reaches the threshold Vthr, the potential is reset to Vreset and the neuron emits a spike, which is sent through the neu- ron’s axon to the target core and neuron. This design implies that each neuron can connect to at most n neurons, which are necessarily in the same core. The initial voltage of each neuron was initialized by drawing randomly and with equal probability from a set of 4 evenly spaced values from Vreset to Vthr. Signal-to-rate transduction The input to the neuromor- phic chip consists of a set of spike trains fed to the neurons of the input layer. To transform the vector signal s into spike trains, we first shifted the signal by ¯s = 3σ, where σ is the standard deviation across all signal components of all pat- terns. The shifted signal was then scaled by a factor νsc cho- sen to ensure moderate output rates in the RCN layer, and the result was linear-rectified to positive values. In short, the input rate νi associated with signal si is νi = νsc[si + ¯s]+, i = 1, . . . , Nin, where [x]+ is x if x > 0, or 0 otherwise. The values νi were then used to generate regular spike trains with fixed inter-spike-interval 1/νi. Basic architecture The circuit is divided in two functional groups, or layers, each of which comprises several cores. The first functional group is the RCN layer, which computes the random nonlinear expansion in Eq. [ 1 ]. The second func- tional group computes the C-class discriminant y = Jx. The output of the classifier is just argmaxj yj, where j runs over the C possible categories. The argmax operation was not com- puted by the chip, but was determined off-line by comparing the accumulated spike counts across all outputs. In the fol- lowing, we describe the implementation of the two layers in more detail. RCN layer We first set the dimensionality of the input to the number of available axons per core, i.e., Nin = n = 256. A convenient choice for W is a n× N matrix where each column is vector of zeros except for exactly m < n nonzero entries, which are randomly placed and take a fixed integer value w. We took m = 26, which corresponds to a connectivity level of around 0.1. Lowering the connectivity has the advantage of decreasing energy costs by reducing the number of total spikes and active synapses, without impacting the classifica- tion performance. The random expansion was mapped in the chip by splitting the matrix WT into (cid:100)N/n(cid:101) submatrices of size n× n, and using each submatrix as the (boolean) connec- tivity matrix Bij of a core. With this arrangement, each of the N neurons distributed among the (cid:100)N/n(cid:101) cores receives a sparse and random linear combination of signals. Specifically, the average current re- ceived by each RCN is n(cid:88) hi = Wjiνj, i = 1, . . . , N. j=1 8 A zero-th order approximation of the firing rate of a gen- eral vlsi neuron receiving a current hi is ri = [hi − β]+ Vthr − Vreset , [ 3 ] where Vthr is the threshold for spike emission and Vreset is the reset potential [12]. We chose the parameters w and β to meet two criteria. First, we required the fraction of RCNs showing any firing activity (i.e., the coding level f ) to be around 0.25. This coding level is a good compromise between the need for dis- crimination and generalization, and it keeps finite-size effects at bay [9]. Second, we required the distribution of activities across active RCNs to be sufficiently wide. Otherwise the in- formation carried by the spiking activity of the RCNs is too imprecise to discriminate among patterns. All the cores in the RCN layer receive exactly the same n-dimensional input signal. Readout The readout matrix J was trained offline and mapped to the chip architecture as follows. Weight quantization Because the chip can hold only integer-valued synapses, we need to map the set of all compo- nents of J into an appropriate finite set of integers. We started clipping the synaptic weights within the bounds (−4σ, 4σ), where σ is the standard deviation of the sample composed of all the components of J. We then rescaled the weights to a convenient magnitude Jmax = 28 (see below), and rounded the weight values to the nearest integer. Weight assignment The TrueNorth connectivity con- straints dictate that each RCN can project to only one axon, meaning that there are at most n = 256 synaptic contacts available to encode the C = 10 weights, J0i, . . . , J9i associ- ated with the i-th RCN. We allocated 24 contacts per class and per axon (see Fig. 5). Each of these 24 contacts were divided in four groups comprising 6 weights each, with val- ues 1, 2, 4,−1,−2,−4. This allowed us to represent any inte- ger weight from −28 to 28 (each of the 4 groups encodes a maximum weight of 7, sign aside). To distribute any weight value w across the available synaptic contacts, we decom- posed w in a sum of four terms, given by the integer divi- sion of w by 4 with the remainder spread evenly across terms (Ex: 19 = 4 + 5 + 5 + 5). Each of such values was assigned to one group, represented in base 2, and mapped to a pat- tern of active-inactive synapses according to the weight as- sociated with each axon-dendrite intersection. Positive and negative weights, as well as strong and week weights, were balanced along a dendrite by changing the sign and order of the weights in the crossbar (see alternating colors and satura- tions in Fig. 5). Negative threshold For the readout to work properly, the firing activity of readout neurons must be proportional to the linear sum of the inputs from the RCNs. This requires neu- rons to operate in the linear regime of their dynamic range, a regime that can be enforced by lowering the threshold βout of readout neurons. We set βout < 0, which is equivalent to adding a constant positive current to each neuron. If the current-to-rate transduction function were the threshold- linear function of Eq. [ 3 ], the baseline activity induced by this constant current would be βout/(Vthr − Vreset) per read- out neuron. The contribution of this background signal should Figure 5. Implementation of the readout matrix in a core. The diagram represents the first 8 input lines and first 48 dendrites (two output units) of a typical readout core. Under each axon-dendrite contact is a square that indicates the potential synaptic strength at the site: color indicates whether the connection is excitatory (red) or inhibitory (blue), while the saturation level represents the absolute value of the synaptic strength, which can be 1, 2, or 4 (low, medium, and high saturation, respectively). Only the sites marked with a dot are active. The green frame highlights all the synaptic contacts allocated for an arbitrary weight of the readout matrix, in this case J13 = −9, which is decomposed as the 4-term sum −9 = −2 − 3 − 2 − 2 = −0102 − 0112 − 0102 − 0102. Note that in this particular axon the ordering of weights is 20, 21, 22 (rightmost bit is the most significant). be subtracted from the readout outputs if one wants to get the equivalent to Eq. [ 2 ], although the step is unnecessary if one only wishes to compare output magnitudes (as we implicitly do in order to find the maximal output). Support Vector Machines. We trained SVMs to perform mul- ticlass classifications based on a one-vs-all scheme, so that the number of output units coincides with the number of classes (as in the neural classifier). SVMs were evaluated us- ing arc-cosine kernels, whic mimic the computation of large feedforward networks with one or more layers of hidden non- linear units [18]. For our particular architecture, based on one hidden layer built with threshold-linear units, the kernel is k(x, y) = (cid:107)x(cid:107)(cid:107)y(cid:107)J1(θ), where J1(θ) = sin θ + (π − θ) cos θ and θ is the angle between the inputs x and y. We considered three types of SVM. For the standard SVM we used the open library libsvm [44], which we patched to in- clude the arccos kernel. The other two SVMs reduce the num- ber of support vectors without sacrificing performance sub- stantially. One of such algorithms is primalSVC, which selects greedily the basis functions by optimizing the primal objec- tive function [45]. The other method is based on the so-called Cutting-Plane Subspace Pursuit algorithm, which reduces the number of support vectors by using basis functions that, un- like standard SVMs, are not necessarily training vectors [46]. Such method is implemented in the library SVMperf. Unlike the other two classifiers, SVMperf used RBF kernels instead of arccos kernels. Estimation of the IBM chip energy consumptionThe en- ergy consumption of the IBM chip was estimated from the TrueNorth specifications [6]. The total energy consumption comprises the baseline energy (15.9 µW per core), the energy to emit spikes (109 pJ per spike), the energy needed to read active synapses (10.7 pJ per active synapse), and the energy necessary to update membrane potentials (1.2 pJ per neuron). We ignored the input-output energy needed to transmit spikes off chip and receive spikes on chip. These numbers provide a reasonable estimate of the energy consumption of systems with a conservative supply voltage of 0.775 V; most chips op- erate near or below this estimate. For a setup with 214 RCNs, 26 dendrites per class, and 10 classes, the power was about 2.08 mW, 95% of which corresponds to the baseline power. Scaling of the energy with the number of classes The estimation was based on the energy cost of the simulated clas- sifications of the MNIST dataset, and extrapolated to the de- signs required by an increasing number of classes. As the number of classes C increases, so does the number of readout neurons necessary to perform a classification and, therefore, so does the required number of readout cores. Specifically, if we assign sc synaptic contacts per axon and per class, we will need a total of scC output lines. These output lines need to be connected to all the N neurons through the input lines of the readout cores . Because each readout core can accomodate 256 output lines, connected to 256 input lines, the total num- ber of readout cores will be (cid:100)N/256(cid:101)(cid:100)scC/256(cid:101) ((cid:100)·(cid:101) indicates the ceiling function). In principle the number of RCN cores will be simply (cid:100)N/256(cid:101). However, each RCN should project to (cid:100)scC/256(cid:101) cores, which implies that each RCN core must be cloned (cid:100)scC/256(cid:101) times due to the fan-out constraint—each RCN can project to only one core. The total number of cores is therefore Ncores = 2(cid:100)N/256(cid:101)(cid:100)scC/256(cid:101), where the factor 2 accounts for the contributions of both the readout and the RCN cores. The total number of spikes emitted was estimated from the reference value we got from the chip simulation (for 10 classes, N = 214, sc = 24, and 500 ms of classification time), scaled appropriately for the new Ncores. More concretely, if we denote by n0 sp the number of spikes emitted during our reference simulation, the number of emitted spikes in a gen- sp(cid:100)scC/256(cid:101)(T /500)(N/214), where T is eral case is nsp = n0 the duration of the simulation in milliseconds. We chose this duration to be T = 108 ms, which is the average classification time of the chip implementation the MNIST dataset, when the spike difference is 80 spikes and which yields only 0.1% less in performance than in the fixed-duration case (97.2% vs 97.3%). With T and the estimated values of Ncores and Figure 6. Simulation of a digital support vector machine. a Number of operations (black circles, left ordinate) and runtime (blue dots, right ordi- nate) required by a digital SVM to classify 10 test patterns from the MNIST dataset, as a function of the number of support vectors. The SVM performance was estimated with a simulator of the Intel i7 processor. b Energy consumption associated to the datapoints shown in a (squares). The straight line is a least-square fit. 9 fromRCN12345678J13=−9+0+1025005000750002.5·1085·10800.050.1numberofSVsnumberofoperationsrunningtime(s)ab025005000750000.20.4numberofSVsenergyconsumption(J)0.053mJperSV(10classifications) Figure 7. Performance versus energy consumption at a fixed classification time. Panels are like in Figs. 3b,d, classification time is now fixed at 500 ms, rather than determined by a stopping criterion. nsp, it is straightforward to compute the energy consumption according to the values given in the previous paragraph. Energy consumption in von Neumann digital machines Configuration The runtime and power of microprocessors with von Neumann architectures were estimated with the re- cently developed simulators GEM5 (gem5.opt 2.0) [47] and McPAT (ver. 1.2) [48]. For the estimation we used an architec- ture configuration similar to that of the recent Intel CoreTM i7 processors [49], which incorporate state-of-the-art CMOS technology. Specifically, we used an x86 64, O3, single core architecture at 2.66 GHz clock frequency, with 32KB 8-way L1-i and 32KB 8-way L1-d caches, 256KB 8-way L2 cache, 64B cache line size, and 8GB DDR3 1600 DRAM. Channel length was 22 nm, HP type, using long channel if appropriate. VDD was 0.9V, so slightly higher than the 0.775 V used for the IBM chip. However, could we use the same voltage in Intel i7 simulator, the energy consumption would be lower by a factor (0.775/0.9)2 = 0.74. This 26% reduction would not change the main conclusions about the energy consumption gap between the IBM chip and the conventional von Neumann digital machines, which is 2–3 orders of magnitude. Simulations The benchmark was the test phase of the SVMs, already trained. Simulations showed that a modern microprocessor based on a von-Neumann architecture takes 115.5 ms to evaluate the test set with 8087 SVs, while con- suming 424.6 mJ (DRAM energy consumption not included). When we varied the number of support vectors from 9 to 8087, both the runtime and energy consumption grew propor- tionally to the number of SVs, while the power was roughly constant due to the fixed hardware configuration (see Fig 6). To estimate how the energy used by von Neumann digital SVMs scales with the number of classes, we ran another set of simulations with Intel i7 simulator, this time varying both the number of support vectors and the number of classes in the classification problem. This step was necessary to deter- mine the overhead incurred when we increase the number of output units. For a given number of classes, the energy cost per support vector was estimated from the least-square fit of the energies against the number of support vectors. Mobile processor We also investigated the runtime and energy consumption of a more energy-efficient but slower mobile microprocessor performing the same target workload. The architecture configuration was: ARMv7, O3, single core, 1GHz CPU clock frequency, 32kB 4-way L1i and 32kB 4-way L1d caches, and 128kB 8-way L2 cache, which is similar to the architecture of ARM Cortex-A9 [50]. The technology node (22 nm) and simulators were the same as in the experiment with the microprocessor mimicking Intel Core i7. For the benchmark code with the largest number of SVs, the task re- quired 1.2· 1010 operations that took 6.35 s at a cost of 7.34 J. Discussion on Intel Xeon Phi Massively parallel archi- tectures have gained a significant amount of attention to improve the throughput and power efficiency of the high- performance computing (HPC) technology, in response to the relatively stagnated improvement in clock frequency. The Xeon Phi coprocessor, recently developed by Intel, is one of such efforts [51]. It integrates more than 50 CPU cores to- gether with L1/L2 caches, network-on-chips, GDDR memory controller, and PCIe interface. Each core supports up-to 4- thread in-order operation and the 512b SIMD VPU (Vector processing unit). While the runtime and energy-consumption of the coprocessor are highly dependent on the target work- loads, several recent investigations quantified the performance and energy-efficiency. In the high-performance configuration, the system integrating Xeon and Xeon Phi shows the through- put of 100 Tera floating-point operations (flop) per second, the power consumption of 72.9 kW, marking the energy efficiency of 0.74 nJ/flop [51]. The classification benchmark codes (with the largest number of SVs) require 0.02235 Gigaflop on the desktop processor configuration similar to Intel Core i7. At a first order approximation, therefore, the Xeon and Xeon Phi-based system takes 0.2235 µs and uses 16.5 mJ per clas- sification. This energy consumption seems significantly lower than the one of the Intel Core i7, and very close to its lower bound, which is approximately 3 mJ. However, one should keep in mind that the energy is grossly underestimated, as not only we ignored the energy needed for the RAM, but we also neglected the cost of the non floating point operations, which are approximately twice as many as the floating point operations. For all these reasons it is difficult to compare the energy consumption for the Xean Phi to the Intel Core i7. In any case, even for our very conservative energy consump- tion estimate, the IBM chip remains significantly more energy efficient. 10 a60801000.1110100performance(%)averageenergyconsumption(mJ)networkchiplibsvmSVCperfprimalSVC2550750.1110performance(%)averageenergyconsumption(mJ)networkchipblibsvmSVCperfprimalSVCMNISTMNIST-back-image 1. Mead C (1989) Analog VLSI implementation of neural systems (Addison Wesley Publishing 28. Hasler J, Marr B (2013) Finding a roadmap to achieve large neuromorphic hardware systems. Company). Indiveri G, et al. (2011) Neuromorphic silicon neuron circuits. Front. Neurosci. 5. 2. 3. Livi P, Indiveri G (2009) A current-mode conductance-based silicon neuron for address-event neuromorphic systems pp 2898–2901. Front. Neurosci. 7. 29. Arthur JV, Boahen K (2011) Silicon-neuron design: A dynamical systems approach. Circuits and Systems I: Regular Papers, IEEE Transactions on 58:1034–1043. 30. Han J, Orshansky M (2013) Approximate computing: An emerging paradigm for energy- 4. Rangan V, Ghosh A, Aparin V, Cauwenberghs G (2010) A subthreshold aVLSI implementation efficient design (IEEE), pp 1–6. of the Izhikevich simple neuron model pp 4164–4167. 31. Krizhevsky A, Sutskever I, Hinton GE (2012) ImageNet classification with deep convolutional 5. Chicca E, Stefanini F, Indiveri G (2014) Neuromorphic electronic circuits for building au- neural networks pp 1097–1105. tonomous cognitive systems. Proceedings of the IEEE PP:1–22. 6. Merolla PA, et al. (2014) A million spiking-neuron integrated circuit with a scalable commu- nication network and interface. Science 345:668–673. 7. Jaeger H, Haas H (2004) Harnessing nonlinearity: predicting chaotic systems and saving energy in wireless communication. Science 304:78–80. 8. Buonomano DV, Maass W (2009) State-dependent computations: spatiotemporal processing in cortical networks. Nat Rev Neurosci 10:113–125. 9. Barak O, Rigotti M, Fusi S (2013) The sparseness of mixed selectivity neurons controls the generalization–discrimination trade-off. J. Neurosci. 33:3844–3856. 10. Arthur JV, et al. (2012) Building block of a programmable neuromorphic substrate: A digital neurosynaptic core (IEEE), pp 1–8. 11. Eliasmith C, et al. (2012) A large-scale model of the functioning brain. Science 338:1202–1205. 12. Fusi S, Mattia M (1999) Collective behavior of networks with linear (VLSI) integrate-and-fire neurons. Neural Comput. 11:633–652. 13. Sompolinsky H (1986) Neural networks with non-linear synapses and static noise. Phys. Rev. A 34:2571. 14. Amit DJ, Fusi S (1994) Learning in neural networks with material synapses. Neural Compu- tation 6:957–982. 15. Fusi S (2002) Hebbian spike-driven synaptic plasticity for learning patterns of mean firing rates. Biological cybernetics 87:459–470. 16. Garey MR, Johnson DS (1979) Computers and intractability: a guide to NP-completeness (WH Freeman New York). 17. Goodfellow I, Warde-farley D, Mirza M, Courville A, Bengio Y (2013) Maxout Networks pp 1319–1327. 32. Deng L, Hinton G, Kingsbury B (2013) New types of deep neural network learning for speech recognition and related applications: An overview (IEEE), pp 8599–8603. 33. Mitra S, Fusi S, Indiveri G (2009) Real-time classification of complex patterns using spike- based learning in neuromorphic vlsi. Biomedical Circuits and Systems, IEEE Transactions on 3:32–42. 34. Giulioni M, et al. (2011) Robust working memory in an asynchronously spiking neural network realized with neuromorphic vlsi. Frontiers in neuroscience 5. 35. Arthur J, Boahen K (2006) Learning in silicon: timing is everything. Advances in neural information processing systems 18:75. 36. LeCun Y, Bottou L, Bengio Y, Haffner P (1998) Gradient-based learning applied to document recognition. Proc. IEEE 86:2278–2324. 37. Larochelle H, Erhan D, Courville A, Bergstra J, Bengio Y (2007) An empirical evaluation of deep architectures on problems with many factors of variation, ICML ’07 (ACM, New York, NY, USA), pp 473–480. 38. Amit Y, Geman D (1997) Shape quantization and recognition with randomized trees. Neural Comput. 9:1545–1588. 39. Amit Y (2002) 2D Object Detection and Recognition: Models, Algorithms, and Networks (MIT Press). 40. Raiko T, Valpola H, LeCun Y (2012) Deep Learning Made Easier by Linear Transformations in Perceptrons Vol. 22, pp 924–932. 41. Huang GB, Zhu QY, Siew CK (2006) Extreme learning machine: theory and applications. Neurocomputing 70:489–501. 42. Tapson J, van Schaik A (2013) Learning the pseudoinverse solution to network weights. Neural Netw. 45:94–100. 43. Merolla P, et al. (2011) A digital neurosynaptic core using embedded crossbar memory with 18. Cho Y, Saul LK (2010) Large-margin classification in infinite neural networks. Neural Comput. 45pJ per spike in 45nm (IEEE), pp 1–4. 22:2678–2697. 19. Sohn K, Zhou G, Lee C, Lee H (2013) Learning and Selecting Features Jointly with Point-wise Gated Boltzmann Machines pp 217–225. 44. Chang CC, Lin CJ (2011) LIBSVM: A library for support vector machines. ACM T. Intel. Sys. Techn. 2:27:1–27:27 Software available at http://www.csie.ntu.edu.tw/~cjlin/libsvm. 45. Keerthi SS, Chapelle O, DeCoste D (2006) Building support vector machines with reduced 20. Lan G, Sartori P, Neumann S, Sourjik V, Tu Y (2012) The energy-speed-accuracy trade-off in classifier complexity. J. Mach. Learn. Res. 7:1493–1515. sensory adaptation. Nature physics 8:422–428. 46. Joachims T, Yu CN (2009) Sparse kernel SVMs via cutting-plane training. Mach. Learn. 21. Tang H, et al. (2014) Spatiotemporal dynamics underlying object completion in human ventral 76:179–193. visual cortex. Neuron 83:736–748. 47. Binkert N, et al. (2011) The GEM5 simulator. ACM SIGARCH Computer Architecture News 22. Boser BE, Guyon IM, Vapnik VN (1992) A training algorithm for optimal margin classifiers 39:1–7. (ACM), pp 144–152. 23. Cortes C, Vapnik V (1995) Support-vector networks. Mach. Learn. 20:273–297. 24. Vapnik V, Golowich SE, Smola A (1997) Support vector method for function approximation, 48. Li S, et al. (2009) McPAT: an integrated power, area, and timing modeling framework for multicore and manycore architectures (IEEE), pp 469–480. 49. (2014) Intel core i7 processor., Technical report http://www.intel.com/content/www/us/en/ regression estimation, and signal processing pp 281–287. processors/core- i7- processor.html. 25. Steinwart I, Christmann A (2008) Support vector machines (Springer). 26. Rahimi A, Recht B (2008) Weighted sums of random kitchen sinks: Replacing minimization with randomization in learning pp 1313–1320. 27. Le Q, Sarl´os T, Smola A (2013) Fastfoodapproximating kernel expansions in loglinear time. 50. (2014) Arm cortex-a9., Technical report http://www.arm.com/products/processors/cortex- a/ cortex- a9.php. 51. Chrysos G, Engineer SP (2012) Intel Xeon Phi coprocessor (codename knights corner). 11
1701.07775
1
1701
2017-01-26T16:59:20
A Forward Model at Purkinje Cell Synapses Facilitates Cerebellar Anticipatory Control
[ "q-bio.NC", "eess.SY", "math.OC" ]
How does our motor system solve the problem of anticipatory control in spite of a wide spectrum of response dynamics from different musculo-skeletal systems, transport delays as well as response latencies throughout the central nervous system? To a great extent, our highly-skilled motor responses are a result of a reactive feedback system, originating in the brain-stem and spinal cord, combined with a feed-forward anticipatory system, that is adaptively fine-tuned by sensory experience and originates in the cerebellum. Based on that interaction we design the counterfactual predictive control (CFPC) architecture, an anticipatory adaptive motor control scheme in which a feed-forward module, based on the cerebellum, steers an error feedback controller with counterfactual error signals. Those are signals that trigger reactions as actual errors would, but that do not code for any current or forthcoming errors. In order to determine the optimal learning strategy, we derive a novel learning rule for the feed-forward module that involves an eligibility trace and operates at the synaptic level. In particular, our eligibility trace provides a mechanism beyond co-incidence detection in that it convolves a history of prior synaptic inputs with error signals. In the context of cerebellar physiology, this solution implies that Purkinje cell synapses should generate eligibility traces using a forward model of the system being controlled. From an engineering perspective, CFPC provides a general-purpose anticipatory control architecture equipped with a learning rule that exploits the full dynamics of the closed-loop system.
q-bio.NC
q-bio
A Forward Model at Purkinje Cell Synapses Facilitates Cerebellar Anticipatory Control Ivan Herreros-Alonso SPECS lab Xerxes D. Arsiwalla SPECS lab Universitat Pompeu Fabra Universitat Pompeu Fabra Barcelona, Spain [email protected] Barcelona, Spain Paul F.M.J. Verschure SPECS, UPF Catalan Institution of Research and Advanced Studies (ICREA) Barcelona, Spain Abstract How does our motor system solve the problem of anticipatory control in spite of a wide spectrum of response dynamics from different musculo-skeletal sys- tems, transport delays as well as response latencies throughout the central nervous system? To a great extent, our highly-skilled motor responses are a result of a reactive feedback system, originating in the brain-stem and spinal cord, combined with a feed-forward anticipatory system, that is adaptively fine-tuned by sensory experience and originates in the cerebellum. Based on that interaction we design the counterfactual predictive control (CFPC) architecture, an anticipatory adaptive motor control scheme in which a feed-forward module, based on the cerebellum, steers an error feedback controller with counterfactual error signals. Those are signals that trigger reactions as actual errors would, but that do not code for any cur- rent or forthcoming errors. In order to determine the optimal learning strategy, we derive a novel learning rule for the feed-forward module that involves an eligibility trace and operates at the synaptic level. In particular, our eligibility trace provides a mechanism beyond co-incidence detection in that it convolves a history of prior synaptic inputs with error signals. In the context of cerebellar physiology, this solution implies that Purkinje cell synapses should generate eligibility traces using a forward model of the system being controlled. From an engineering perspective, CFPC provides a general-purpose anticipatory control architecture equipped with a learning rule that exploits the full dynamics of the closed-loop system. 1 Introduction Learning and anticipation are central features of cerebellar computation and function (Bastian, 2006): the cerebellum learns from experience and is able to anticipate events, thereby complementing a reactive feedback control by an anticipatory feed-forward one (Hofstoetter et al., 2002; Herreros and Verschure, 2013). This interpretation is based on a series of anticipatory motor behaviors that originate in the cerebellum. For instance, anticipation is a crucial component of acquired behavior in eye-blink conditioning (Gormezano et al., 1983), a trial by trial learning protocol where an initially neutral stimulus such as a tone or a light (the conditioning stimulus, CS) is followed, after a fixed delay, by a noxious one, such as an air puff to the eye (the unconditioned stimulus, US). During early trials, a protective unconditioned response (UR), a blink, occurs reflexively in a feedback manner following the US. After training though, a well-timed anticipatory blink (the conditioned response, CR) precedes the US. Thus, learning results in the (partial) transference from an initial feedback action to an anticipatory (or predictive) feed-forward one. Similar responses occur during anticipatory postural adjustments, which are postural changes that precede voluntary motor movements, such as raising an arm while standing (Massion, 1992). The goal of these anticipatory adjustments is to counteract the postural and equilibrium disturbances that voluntary movements introduce. These 30th Conference on Neural Information Processing Systems (NIPS 2016), Barcelona, Spain. behaviors can be seen as feedback reactions to events that after learning have been transferred to feed-forward actions anticipating the predicted events. Anticipatory feed-forward control can yield high performance gains over feedback control whenever the feedback loop exhibits transmission (or transport) delays (Jordan, 1996). However, even if a plant has negligible transmission delays, it may still have sizable inertial latencies. For example, if we apply a force to a visco-elastic plant, its peak velocity will be achieved after a certain delay; i.e. the velocity itself will lag the force. An efficient way to counteract this lag will be to apply forces anticipating changes in the desired velocity. That is, anticipation can be beneficial even when one can act instantaneously on the plant. Given that, here we address two questions: what is the optimal strategy to learn anticipatory actions in a cerebellar-based architecture? and how could it be implemented in the cerebellum? To answer that we design the counterfactual predictive control (CFPC) scheme, a cerebellar-based adaptive-anticipatory control architecture that learns to anticipate performance errors from experience. The CFPC scheme is motivated from neuro-anatomy and physiology of eye-blink conditioning. It includes a reactive controller, which is an output-error feedback controller that models brain stem reflexes actuating on eyelid muscles, and a feed-forward adaptive component that models the cerebellum and learns to associate its inputs with the error signals driving the reactive controller. With CFPC we propose a generic scheme in which a feed-forward module enhances the performance of a reactive error feedback controller steering it with signals that facilitate anticipation, namely, with counterfactual errors. However, within CFPC, even if these counterfactual errors that enable predictive control are learned based on past errors in behavior, they do not reflect any current or forthcoming error in the ongoing behavior. In addition to eye-blink conditioning and postural adjustments, the interaction between reactive and cerebellar-dependent acquired anticipatory behavior has also been studied in paradigms such as visually-guided smooth pursuit eye movements (Lisberger, 1987). All these paradigms can be abstracted as tasks in which the same predictive stimuli and disturbance or reference signal are repeatedly experienced. In accordance to that, we operate our control scheme in trial-by-trial (batch) mode. With that, we derive a learning rule for anticipatory control that modifies the well-known least-mean-squares/Widrow-Hoff rule with an eligibility trace. More specifically, our model predicts that to facilitate learning, parallel fibers to Purkinje cell synapses implement a forward model that generates an eligibility trace. Finally, to stress that CFPC is not specific to eye-blink conditioning, we demonstrate its application with a smooth pursuit task. 2 Methods 2.1 Cerebellar Model Figure 1: Anatomical scheme of a Cerebellar Purkinje cell. The xj denote parallel fiber inputs to Purkinje synapses (in red) with weights wj. o denotes the output of the Purkinje cell. The error signal e, through the climbing fibers (in green), modulates synaptic weights. We follow the simplifying approach of modeling the cerebellum as a linear adaptive filter, while focusing on computations at the level of the Purkinje cells, which are the main output cells of the cerebellar cortex (Fujita, 1982; Dean et al., 2010). Over the mossy fibers, the cerebellum receives a wide range of inputs. Those inputs reach Purkinke cells via parallel fibers (Fig. 1), that cross 2 xj x1 xN oew1wjwN dendritic trees of Purkinje cells in a ratio of up to 1.5 × 106 parallel fiber synapses per cell (Eccles et al., 1967). We denote the signal carried by a particular fiber as xj, j ∈ [1, G], with G equal to the total number of inputs fibers. These inputs from the mossy/parallel fiber pathway carry contextual information (interoceptive or exteroceptive) that allows the Purkinje cell to generate a functional output. We refer to these inputs as cortical bases, indicating that they are localized at the cerebellar cortex and that they provide a repertoire of states and inputs that the cerebellum combines to generate its output o. As we will develop a discrete time analysis of the system, we use n to indicate time (or time-step). The output of the cerebellum at any time point n results from a weighted sum of those cortical bases. wj indicates the weight or synaptic efficacy associated with the fiber j. Thus, we (cid:124) (where the transpose, (cid:124), indicates have x[n] = [x1[n], . . . , xG[n]] that x[n] and w[n] are column vectors) containing the set of inputs and synaptic weights at time n, respectively, which determine the output of the cerebellum according to (cid:124) and w[n] = [w1[n], . . . , wG[n]] o[n] = x[n] (cid:124) w[n] (1) The adaptive feed-forward control of the cerebellum stems from updating the weights according to a rule of the form ∆wj[n + 1] = f (xj[n], . . . , xj[1], e[n], Θ) (2) where Θ denotes global parameters of the learning rule; xj[n], . . . , xj[1], the history of its pre- synaptic inputs of synapse j; and e[n], an error signal that is the same for all synapses, corresponding to the difference between the desired, r, and the actual output, y, of the controlled plant. Note that in drawing an analogy with the eye-blink conditioning paradigm, we use the simplifying convention of considering the noxious stimulus (the air-puff) as a reference, r, that indicates that the eyelids should close; the closure of the eyelid as the output of the plant, y; and the sensory response to the noxious stimulus as an error, e, that encodes the difference between the desired, r, and the actual eyelid closures, y. Given this, we advance a new learning rule, f, that achieves optimal performance in the context of eye-blink conditioning and other cerebellar learning paradigms. 2.2 Cerebellar Control Architecture Figure 2: Neuroanatomy of eye-blink conditioning and the CFPC architecture. Left: Mapping of signals to anatomical structures in eye-blink conditioning (De Zeeuw and Yeo, 2005); regular arrows indicate external inputs and outputs, arrows with inverted heads indicate neural pathways. Right: CFPC architecture. Note that the feedback controller, C, and the feed-forward module, F F , belong to the control architecture, while the plant, P , denotes an object controlled. Other abbreviations: r, reference signal; y, plant's output; e, output error; x, basis signals; o, feed-forward signal; and u, motor command. We embed the adaptive filter cerebellar module in a layered control architecture, namely the CFPC architecture, based on the interaction between brain stem motor nuclei driving motor reflexes and the cerebellum, such as the one established between the cerebellar microcircuit responsible for conditioned responses and the brain stem reflex circuitry that produces unconditioned eye-blinks (Hesslow and Yeo, 2002) (Fig. 2 left). Note that in our interpretation of this anatomy we assume that cerebellar output, o, feeds the lower reflex controller (Fig. 2 right). Put in control theory terms, within the CFPC scheme an adaptive feed-forward layer supplements a negative feedback controller steering it with feed-forward signals. 3 +-US(airpu(cid:31))[r]Eyelids(Blink)[P][y]Facialnucleus[C]Trigeminalnucleus[e][e]CS(Context, e.g.: sound, light)[u]Cerebellum (cortex and nuclei) and Inferior olive [FF][x]Pons[o]FFxoreCuPy+-+ADAPTIVE(cid:31)ANTICIPATORY(cid:30)FEED(cid:31)FORWARD(cid:29) LAYER REACTIVE (cid:30)FEEDBACK(cid:29) LAYERFEEDBACK CLOSED(cid:31)LOOP SYSTEM Our architecture uses a single-input single-output negative-feedback controller. The controller receives as input the output error e = r − y. For the derivation of the learning algorithm, we assume that both plant and controller are linear and time-invariant (LTI) systems. Importantly, the feedback controller and the plant form a reactive closed-loop system, that mathematically can be seen as a system that maps the reference, r, into the plant's output, y. A feed-forward layer that contains the above-mentioned cerebellar model provides the negative feedback controller with an additional input signal, o. We refer to o as a counter-factual error signal, since although it mechanistically drives the negative feedback controller analogously to an error signal it is not an actual error. The counterfactual error is generated by the feed-forward module that receives an output error, e, as its teaching signal. Notably, from the point of view of the reactive layer closed-loop system, o can also be interpreted as a signal that offsets r. In other words, even if r remains the reference that sets the target of behavior, r + o functions as the effective reference that drives the closed-loop system. 3 Results 3.1 Derivation of the gradient descent update rule for the cerebellar control architecture We apply the CFPC architecture defined in the previous section to a task that consists in following a finite reference signal r ∈ RN that is repeated trial-by-trial. To analyze this system, we use the discrete time formalism and assume that all components are linear time-invariant (LTI). Given this, both reactive controller and plant can be lumped together into a closed-loop dynamical system, that can be described with the dynamics A, input B, measurement C and feed-through D matrices. In general, these matrices describe how the state of a dynamical system autonomously evolves with time, A; how inputs affect system states, B; how states are mapped into outputs, C; and how inputs instantaneously affect the system's output D (Astrom and Murray, 2012). As we consider a reference of a finite length N, we can construct the N-by-N transfer matrix T as follows (Boyd, 2008)  T = D CB CAB 0 D CB 0 0 D ... ... CAN−2B CAN−3B CAN−4B . . . D ... ...  0 0 0 . . . . . . . . . ... With this transfer matrix we can map any given reference r into an output yr using yr = T r, obtaining what would have been the complete output trajectory of the plant on an entirely feedback-driven trial. Note that the first column of T contains the impulse response curve of the closed-loop system, while the rest of the columns are obtained shifting that impulse response down. Therefore, we can build the transfer matrix T either in a model-based manner, deriving the state-space characterization of the closed-loop system, or in measurement-based manner, measuring the impulse response curve. Additionally, note that (I − T )r yields the error of the feedback control in following the reference, a signal which we denote with e0. Let o ∈ RN be the entire feed-forward signal for a given trial. Given commutativity, we can consider that from the point of view of the closed-loop system o is added directly to the reference r, (Fig. 2 right). In that case, we can use y = T (r + o) to obtain the output of the closed-loop system when it is driven by both the reference and the feed-forward signal. The feed-forward module only outputs linear combinations of a set of bases. Let X ∈ RN×G be a matrix with the content of the G bases during all the N time steps of a trial. The feed-forward signal becomes o = Xw, where w ∈ RG contains the mixing weights. Hence, the output of the plant given a particular w becomes y = T (r + Xw). We implement learning as the process of adjusting the weights w of the feed-forward module in a trial-by-trial manner. At each trial the same reference signal, r, and bases, X, are repeated. Through learning we want to converge to the optimal weight vector w∗ defined as (r − T (r + Xw)) (r − T (r + Xw)) w∗ = arg min c(w) = arg min e = arg min (3) (cid:124) (cid:124) e 1 2 1 2 w w w where c indicates the objective function to minimize, namely the L2 norm or sum of squared errors. With the substitution X = T X and using e0 = (I − T )r, the minimization problem can be cast as a 4 canonical linear least-squares problem: w∗ = arg min w 1 2 (cid:124) (e0 − Xw) (e0 − Xw) (4) One the one hand, this allows to directly find the least squares solution for w∗, that is, w∗ = X†e0, where † denotes the Moore-Penrose pseudo-inverse. On the other hand, and more interestingly, with w[k] being the weights at trial k and having e[k] = e0 − Xw[k], we can obtain the gradient of the error function at trial k with relation to w as follows: ∇wc = − X (cid:124) e[k] = −X (cid:124)T (cid:124) e[k] Thus, setting η as a properly scaled learning rate (the only global parameter Θ of the rule), we can derive the following gradient descent strategy for the update of the weights between trials: w[k + 1] = w[k] + ηX (cid:124)T (cid:124) e[k] (5) This solves for the learning rule f in eq. 2. Note that f is consistent with both the cerebellar anatomy (Fig. 2left) and the control architecture (Fig. 2right) in that the feed-forward module/cerebellum only requires two signals to update its weights/synaptic efficacies: the basis inputs, X, and error signal, e. 3.2 T (cid:124) facilitates a synaptic eligibility trace The standard least mean squares (LMS) rule (also known as Widrow-Hoff or decorrelation learning rule) can be represented in its batch version as w[k + 1] = w[k] + ηX e[k]. Hence, the only difference between the batch LMS rule and the one we have derived is the insertion of the matrix factor T (cid:124). Now we will show how this factor acts as a filter that computes an eligibility trace at each weight/synapse. Note that the update of a single weight, according Eq. 5 becomes (cid:124) (6) where xj contains the sequence of values of the cortical basis j during the entire trial. This can be rewritten as e[k] (cid:124) jT (cid:124) wj[k + 1] = wj[k] + ηx with hj ≡ T xj. The above inner product can be expressed as a sum of scalar products (cid:124) wj[k + 1] = wj[k] + ηh j e[k] (7) N(cid:88) wj[k + 1] = wj[k] + η hj[n]e[k, n] (8) n=1 where n indexes the within trial time-step. Note that e[k] in Eq. 7 refers to the whole error signal at trial k whereas e[k, n] in Eq. 8 refers to the error value in the n-th time-step of the trial k. It is now clear that each hj[n] weighs how much an error arriving at time n should modify the weight wj, which is precisely the role of an eligibility trace. Note that since T contains in its columns/rows shifted repetitions of the impulse response curve of the closed-loop system, the eligibility trace codes at any time n, the convolution of the sequence of previous inputs with the impulse-response curve of the reactive layer closed-loop. Indeed, in each synapse, the eligibility trace is generated by a forward model of the closed-loop system that is exclusively driven by the basis signal. Consequently, our main result is that by deriving a gradient descent algorithm for the CFPC cerebellar control architecture we have obtained an exact definition of the suitable eligibility trace. That definition guarantees that the set of weights/synaptic efficacies are updated in a locally optimal manner in the weights' space. 3.3 On-line gradient descent algorithm The trial-by-trial formulation above allowed for a straightforward derivation of the (batch) gradient descent algorithm. As it lumped together all computations occurring in a same trial, it accounted for time within the trial implicitly rather than explicitly: one-dimensional time-signals were mapped onto points in a high-dimensional space. However, after having established the gradient descent algorithm, we can implement the same rule in an on-line manner, dropping the repetitiveness assumption inherent to trial-by-trial learning and performing all computations locally in time. Each weight/synapse must 5 have a process associated to it that outputs the eligibility trace. That process passes the incoming (unweighted) basis signal through a (forward) model of the closed-loop as follows: sj[n + 1] = Asj[n] + Bxj[n] hj[n] = Csj[n] + Dxj[n] where matrices A, B, C and D refer to the closed-loop system (they are the same matrices that we used to define the transfer matrix T ), and sj[n] is the state vector of the forward model of the synapse j at time-step n. In practice, each "synaptic" forward model computes what would have been the effect of having driven the closed-loop system with each basis signal alone. Given the superposition principle, the outcome of that computation can also be interpreted as saying that hj[n] indicates what would have been the displacement over the current output of the plant, y[n], achieved feeding the closed-loop system with the basis signal xj. The process of weight update is completed as follows: (9) At each time step n, the error signal e[n] is multiplied by the current value of the eligibility trace hj[n], scaled by the learning rate η, and subtracted to the current weight wj[n]. Therefore whereas the contribution of each basis to the output of the adaptive filter depends only on its current value and weight, the change in weight depends on the current and past values passed through a forward model of the closed-loop dynamics. wj[n + 1] = wj[n] + ηhj[n]e[n] 3.4 Simulation of a visually-guided smooth pursuit task We demonstrate the CFPC approach in an example of a visual smooth pursuit task in which the eyes have to track a target moving on a screen. Even though the simulation does not capture all the complexity of a smooth pursuit task, it illustrates our anticipatory control strategy. We model the plant (eye and ocular muscles) with a two-dimensional linear filter that maps motor commands into angular positions. Our model is an extension of the model in (Porrill and Dean, 2007), even though in that work the plant was considered in the context of the vestibulo-ocular reflex. In particular, we use a chain of two leaky integrators: a slow integrator with a relaxation constant of 100 ms drives the eyes back to the rest position; the second integrator, with a fast time constant of 3 ms ensures that the change in position does not occur instantaneously. To this basic plant, we add a reactive control layer modeled as a proportional-integral (PI) error-feedback controller, with proportional gain kp and integral gain ki. The control loop includes a 50 ms delay in the error feedback, to account for both the actuation and the sensing latency. We choose gains such that reactive tracking lags the target by approximately 100 ms. This gives kp = 20 and ki = 100. To complete the anticipatory and adaptive control architecture, the closed-loop system is supplemented by the feed-forward module. Figure 3: Behavior of the system. Left: Reference (r) and output of the system before (y[1]) and after learning (y[50]). Right: Error before e[1] and after learning e[50] and output acquired by cerebellar/feed-forward component (o[50]) The architecture implementing the forward model-based gradient descent algorithm is applied to a task structured in trials of 2.5 sec duration. Within each trial, a target remains still at the center of the visual scene for a duration 0.5 sec, next it moves rightwards for 0.5 sec with constant velocity, remains still for 0.5 sec and repeats the sequence of movements in reverse, returning to the center. The cerebellar component receives 20 Gaussian basis signals (X) whose receptive fields are defined in the temporal domain, relative to trial onset, with a width (standard-deviation) of 50 ms and spaced by 100 ms. The whole system is simulated using a 1 ms time-step. To construct the matrix T we computed closed-loop system impulse response. 6 00.511.522.500.20.40.60.81time (s)angular position (a.u.) ry[1]y[50]00.511.522.5−0.100.10.2time (s)angular position (a.u.) e[1]e[50]o[50] At the first trial, before any learning, the output of the plant lags the reference signal by approximately 100 ms converging to the position only when the target remains still for about 300 ms (Fig. 3 left). As a result of learning, the plant's behavior shifts from a reactive to an anticipatory mode, being able to track the reference without any delay. Indeed, the error that is sizable during the target displacement before learning, almost completely disappears by the 50th trial (Fig. 3 right). That cancellation results from learning the weights that generate a feed-forward predictive signal that leads the changes in the reference signal (onsets and offsets of target movements) by approximately 100 ms (Fig. 3 right). Indeed, convergence of the algorithm is remarkably fast and by trial 7 it has almost converged to the optimal solution (Fig. 4). Figure 4: Performance achieved with different learning rules. Representative learning curves of the forward model-based eligibility trace gradient descent (FM-ET), the simple Widrow-Hoff (WH) and the Widrow-Hoff algorithm with a delta-eligibility trace matched to error feedback delay (WH+50 ms) or with an eligibility trace exceeding that delay by 20 ms (WH+70 ms). Error is quantified as the relative root mean-squared error (rRMSE), scaled proportionally to the error in the first trial. Error of the optimal solution, obtained with w∗ = (T X)†e0, is indicated with a dashed line. To assess how much our forward-model-based eligibility trace contributes to performance, we test three alternative algorithms. In both cases we employ the same control architecture, changing the plasticity rule such that we either use no eligibility trace, thus implementing the basic Widrow-Hoff learning rule, or use the Widrow-Hoff rule extended with a delta-function eligibility trace that matches the latency of the error feedback (50 ms) or slightly exceeds it (70 ms). Performance with the basic WH model worsens rapidly whereas performance with the WH learning rule using a "pure delay" eligibility trace matched to the transport delay improves but not as fast as with the forward-model- based eligibility trace (Fig. 4). Indeed, in this case, the best strategy for implementing a delayed delta eligibility trace is setting a delay exceeding the transport delay by around 20 ms, thus matching the peak of the impulse response. In that case, the system performs almost as good as with the forward-model eligibility trace (70 ms). This last result implies that, even though the literature usually emphasizes the role of transport delays, eligibility traces also account for response lags due to intrinsic dynamics of the plant. To summarize our results, we have shown with a basic simulation of a visual smooth pursuit task that generating the eligibility trace by means of a forward model ensures convergence to the optimal solution and accelerates learning by guaranteeing that it follows a gradient descent. 4 Discussion In this paper we have introduced a novel formulation of cerebellar anticipatory control, consistent with experimental evidence, in which a forward model has emerged naturally at the level of Purkinje cell synapses. From a machine learning perspective, we have also provided an optimality argument for the derivation of an eligibility trace, a construct that was often thought of in more heuristic terms as a mechanism to bridge time-delays (Barto et al., 1983; Shibata and Schaal, 2001; McKinstry et al., 2006). The first seminal works of cerebellar computational models emphasized its role as an associative memory (Marr, 1969; Albus, 1971). Later, the cerebellum was investigates as a device processing correlated time signals(Fujita, 1982; Kawato et al., 1987; Dean et al., 2010). In this latter framework, 7 0102030405000.20.40.60.81#trialrRMSEWHWH+50msWH+70msFM−ET the use of the computational concept of an eligibility trace emerged as a heuristic construct that allowed to compensate for transmission delays in the circuit(Kettner et al., 1997; Shibata and Schaal, 2001; Porrill and Dean, 2007), which introduced lags in the cross-correlation between signals. Concretely, that was referred to as the problem of delayed error feedback, due to which, by the time an error signal reaches a cell, the synapses accountable for that error are no longer the ones currently active, but those that were active at the time when the motor signals that caused the actual error were generated. This view has however neglected the fact that beyond transport delays, response dynamics of physical plants also influence how past pre-synaptic signals could have related to the current output of the plant. Indeed, for a linear plant, the impulse-response function of the plant provides the complete description of how inputs will drive the system, and as such, integrates transmission delays as well as the dynamics of the plant. Recently, Even though cerebellar microcircuits have been used as models for building control architectures, e.g., the feedback-error learning model (Kawato et al., 1987), our CFPC is novel in that it links the cerebellum to the input of the feedback controller, ensuring that the computational features of the feedback controller are exploited at all times. Within the domain of adaptive control, there are remarkable similarities at the functional level between CFPC and iterative learning control (ILC) (Amann et al., 1996), which is an input design technique for learning optimal control signals in repetitive tasks. The difference between our CFPC and ILC lies in the fact that ILC controllers directly learn a control signal, whereas, the CFPC learns a conterfactual error signal that steers a feedback controller. However the similarity between the two approaches can help for extending CFPC to more complex control tasks. With our CFPC framework, we have modeled the cerebellar system at a very high level of abstraction: we have not included bio-physical constraints underlying neural computations, obviated known anatomical connections such as the cerebellar nucleo-olivary inhibition (Bengtsson and Hesslow, 2006; Herreros and Verschure, 2013) and made simplifications such as collapsing cerebellar cortex and nuclei into the same computational unit. On the one hand, such a choice of high-level abstraction may indeed be beneficial for deriving general-purpose machine learning or adaptive control algorithms. On the other hand, it is remarkable that in spite of this abstraction our framework makes fine-grained predictions at the micro-level of biological processes. Namely, that in a cerebellar microcircuit (Apps and Garwicz, 2005), the response dynamics of secondary messengers (Wang et al., 2000) regulating plasticity of Purkinje cell synapses to parallel fibers must mimic the dynamics of the motor system being controlled by that cerebellar microcircuit. Notably, the logical consequence of this prediction, that different Purkinje cells should display different plasticity rules according to the system that they control, has been validated recording single Purkinje cells in vivo (Suvrathan et al., 2016). In conclusion, we find that a normative interpretation of plasticity rules in Purkinje cell synapses emerges from our systems level CFPC computational architecture. That is, in order to generate optimal eligibility traces, synapses must include a forward model of the controlled subsystem. This conclusion, in the broader picture, suggests that synapses are not merely components of multiplicative gains, but rather the loci of complex dynamic computations that are relevant from a functional perspective, both, in terms of optimizing storage capacity (Benna and Fusi, 2016; Lahiri and Ganguli, 2013) and fine-tuning learning rules to behavioral requirements. Acknowledgments The research leading to these results has received funding from the European Commission's Horizon 2020 socSMC project (socSMC-641321H2020-FETPROACT-2014) and by the European Research Council's CDAC project (ERC-2013-ADG 341196). References Albus, J. S. (1971). A theory of cerebellar function. Mathematical Biosciences, 10(1):25–61. Amann, N., Owens, D. H., and Rogers, E. (1996). Iterative learning control for discrete-time systems with exponential rate of convergence. IEE Proceedings-Control Theory and Applications, 143(2):217–224. Apps, R. and Garwicz, M. (2005). Anatomical and physiological foundations of cerebellar information processing. Nature reviews. Neuroscience, 6(4):297–311. Astrom, K. J. and Murray, R. M. (2012). Feedback Systems: An Introduction for Scientists and Engineers. Princeton university press. 8 Barto, A. G., Sutton, R. S., and Anderson, C. W. (1983). Neuronlike adaptive elements that can solve difficult learning control problems. IEEE transactions on systems, man, and cybernetics, SMC-13(5):834–846. Bastian, A. J. (2006). Learning to predict the future: the cerebellum adapts feedforward movement control. Current Opinion in Neurobiology, 16(6):645–649. Bengtsson, F. and Hesslow, G. (2006). Cerebellar control of the inferior olive. Cerebellum (London, England), 5(1):7–14. Benna, M. K. and Fusi, S. (2016). Computational principles of synaptic memory consolidation. Nature neuroscience. Boyd, S. (2008). Introduction to linear dynamical systems. Online Lecture Notes. De Zeeuw, C. I. and Yeo, C. H. (2005). Time and tide in cerebellar memory formation. Current opinion in neurobiology, 15(6):667–74. Dean, P., Porrill, J., Ekerot, C.-F., and Jörntell, H. (2010). The cerebellar microcircuit as an adaptive filter: experimental and computational evidence. Nature reviews. Neuroscience, 11(1):30–43. Eccles, J., Ito, M., and Szentágothai, J. (1967). The cerebellum as a neuronal machine. Springer Berlin. Fujita, M. (1982). Adaptive filter model of the cerebellum. Biological cybernetics, 45(3):195–206. Gormezano, I., Kehoe, E. J., and Marshall, B. S. (1983). Twenty years of classical conditioning with the rabbit. Herreros, I. and Verschure, P. F. M. J. (2013). Nucleo-olivary inhibition balances the interaction between the reactive and adaptive layers in motor control. Neural Networks, 47:64–71. Hesslow, G. and Yeo, C. H. (2002). The functional anatomy of skeletal conditioning. In A neuroscientist's guide to classical conditioning, pages 86–146. Springer. Hofstoetter, C., Mintz, M., and Verschure, P. F. (2002). The cerebellum in action: a simulation and robotics study. European Journal of Neuroscience, 16(7):1361–1376. Jordan, M. I. (1996). Computational aspects of motor control and motor learning. In Handbook of perception and action, volume 2, pages 71–120. Academic Press. Kawato, M., Furukawa, K., and Suzuki, R. (1987). A hierarchical neural-network model for control and learning of voluntary movement. Biological Cybernetics, 57(3):169–185. Kettner, R. E., Mahamud, S., Leung, H. C., Sitkoff, N., Houk, J. C., Peterson, B. W., and Barto, a. G. (1997). Prediction of complex two-dimensional trajectories by a cerebellar model of smooth pursuit eye movement. Journal of neurophysiology, 77:2115–2130. Lahiri, S. and Ganguli, S. (2013). A memory frontier for complex synapses. In Advances in neural information processing systems, pages 1034–1042. Lisberger, S. (1987). Visual Motion Processing And Sensory-Motor Integration For Smooth Pursuit Eye Movements. Annual Review of Neuroscience, 10(1):97–129. Marr, D. (1969). A theory of cerebellar cortex. The Journal of physiology, 202(2):437–470. Massion, J. (1992). Movement, posture and equilibrium: Interaction and coordination. Progress in Neurobiology, 38(1):35–56. McKinstry, J. L., Edelman, G. M., and Krichmar, J. L. (2006). A cerebellar model for predictive motor control tested in a brain-based device. Proceedings of the National Academy of Sciences of the United States of America, 103(9):3387–3392. Porrill, J. and Dean, P. (2007). Recurrent cerebellar loops simplify adaptive control of redundant and nonlinear motor systems. Neural computation, 19(1):170–193. Shibata, T. and Schaal, S. (2001). Biomimetic smooth pursuit based on fast learning of the target dynamics. In Intelligent Robots and Systems, 2001. Proceedings. 2001 IEEE/RSJ International Conference on, volume 1, pages 278–285. IEEE. Suvrathan, A., Payne, H. L., and Raymond, J. L. (2016). Timing rules for synaptic plasticity matched to behavioral function. Neuron, 92(5):959–967. Wang, S. S.-H., Denk, W., and Häusser, M. (2000). Coincidence detection in single dendritic spines mediated by calcium release. Nature neuroscience, 3(12):1266–1273. 9
1902.05845
1
1902
2019-02-15T15:30:38
Assessing the Level of Autonomic Nervous Activity for Effective Biofeedback Training
[ "q-bio.NC", "physics.med-ph" ]
This paper proposes a prototype of a new biofeedback training based on mathematical models of cardiovascular control. For this purpose we develop a low-cost device that is able to record and process arterial pulse wave via photoplethysmograph and skin temperature on the peripheral part of the arm. A benchmark analysis of our device against a registered cardiovascular measurement system (Biopac MP35 from Biopac Inc., USA) shows that heart rate and skin temperature values delivered by our device are acceptable with a very low residual error (+/-1 beat/min, +/-1 oC). Both measured signals are feed into a mathematical model which estimates the level of activity of both sympathetic and parasympathetic branches of the autonomic nervous system. That information is used jointly with other biological parameters for investigating the stress score of the subject. The data is processed in real-time and continuously displayed to the user for effective biofeedback training. The complete solution was preliminary tested on three volunteers who used the displayed biofeedback information in order to regulate their emotional state successfully during Biofeedback training. They exhibited a significant reduction in stress score compared to three control subjects who did not used our solution during biofeedback training. This supports the benefits of biofeedback training with autonomic nervous tone assessment as effective holistic healing method.
q-bio.NC
q-bio
1 Assessing the Level of Autonomic Nervous Activity for Effective Biofeedback Training Michel KANA Dept. of Biomedical Informatics, Czech Technical University in Prague, Nám. Sítná 3105, 272 01 Kladno, Czech Republic Abstract. This paper proposes a prototype of a new biofeedback training based on mathematical models of cardiovascular control. For this purpose we develop a low- cost device that is able to record and process arterial pulse wave via photoplethysmograph and skin temperature on the peripheral part of the arm. A benchmark analysis of our cardiovascular measurement system (Biopac MP35 from Biopac Inc., USA) shows that heart rate and skin temperature values delivered by our device are acceptable with a very low residual error (± 1 beat/min, ± 1ºC). registered against device a Both measured signals are feed into a mathematical model which estimates the level of activity of both sympathetic and parasympathetic branches of the autonomic nervous system. That jointly with other biological parameters for investigating the stress score of the subject. The data is processed in real-time and continuously displayed to the user for effective biofeedback training. information is used The complete solution was preliminary tested on three volunteers who used the displayed biofeedback information in order to regulate their emotional state successfully during Biofeedback training. They exhibited a significant reduction in stress score compared to three control subjects who did not used our solution during biofeedback training. This supports the benefits of biofeedback training with autonomic nervous tone assessment as effective holistic healing method. Keywords Biofeedback, pulse photoplethysmograph, autonomic nervous tone, cardiovascular control, mathematical modeling. 1. Background to is a Biofeedback influence physiological activities learning process for gaining capability that normally happen unconsciously, such as heart function, breathing and skin temperature. Biofeedback can be effective in combination with other therapies for diverse disorders of psychical and physical function such as anxiety, psychological stress, hypertension, attention disorders, urinary incontinence or depressive disorders [1]. for sensors therapists use Biofeedback recording physiological signals such as skin temperature, muscle activity, heart rate, respiration, skin conductance, or brainwave activity. This stream of information is processed and presented in a simplified graphical or audible form that allows patients to sense and influence changes in their physiological activity in real time. The main limitation of traditional biofeedback is that subjects receive very simple audio and video feedback information [2]. they important We believe that assessing the level of autonomic nervous activity and feeding it back to subject's awareness will improve the effectiveness of biofeedback therapy. Autonomic control centers in the brain monitor and regulate functions. They maintain body temperature and body osmolarity; they control reproductive intake; and function, food influence behavior, emotions and the cardiovascular control center [3]. Assessing autonomic function is usually indirectly done by studying plasma or urinary levels of neurotransmitters, baroreflex sensitivity or heart rate variability [4]. Results are however not always reliable [5]. An attempt for improving the quantitative assessment of autonomic cardiovascular control is to combine the information obtained from the recorded physiological signal with the information derived from mathematical models [6]. Such a method is developed in this paper together with a complete hardware and software solution for effective biofeedback training. Fig. 1. Biofeedback solution based on mathematical models of cardiovascular neural control 2. Materials and Methods As depicted in Fig. 1, our biofeedback method is based on measurement of skin temperature and pulse wave. The temperature is recorded using a thermocouple. The pulse wave is recorded using a photoplethysmograph (PPG). The data from both sensors is processed in a 2 mathematical model which provides information about the balanced state of autonomic nervous system to the user. 2.1 Hardware Development signal is 1000 times amplified in order to acquire a signal in order of volts using the instrumentation amplifier INA128 by Texas Instruments. Power supply is obtained from ±9 volts batteries. Fig. 4. Thermocouple circuit used for temperature measurement (R1=56Ω, R2=10kΩ, R3=1kΩ, R4=500Ω, R5=10kΩ, R6=500Ω, R7=10kΩ, P1=10kΩ; C1=47nF, C2=10pF; U1 INA128, U2 TL074) The Object Oriented Programmable Integrated Circuit (OOPic by Savage Innovations) microcontroller was used for the purpose of 5 volts A/D conversion and RS-232 serial communication with the computer. Fig. 2. Schematic of the hardware prototype the We use the photoplethysmographic method for heart rate measurement [7]. Light can be transmitted through tissues, which are full of small capillaries. They are filled via arterial pulsations during each heart beat. The resulting changes in volume of those vessels modify the absorption, reflection, and scattering of the light. By measuring the level of modification of light source using a photoreceptor, it is possible to measure heart rate. As light source we used an AlGaAs (Aluminum, Gallium, Arsenide) LED, which produces a narrow-band source with a peak spectral emission at a wavelength of 850 nm. As photosensor we used a BPW42 phototransistor which is sensible to the infrared band, from 560 to 980 nm, with peak wavelength sensitivity at 830 nm. As depicted in Fig. 3, the output of a light-emitting diode is altered by tissue absorption and modulates the phototransistor. The DC level is blocked by the capacitor. A noninverting amplifier drives low impedance loads, and provides a gain of 100. Power supply of photoplethysmograph circuit is obtained from 4.5 volts batteries. Fig. 5. Box with boards, power supply and supply switches The final hardware prototype is shown in Fig. 5. We validated our hardware solution against the well-known Biopac MP35 system, developed by Biopac Systems Inc., which is a registered system for physiological data acquisition and analysis. Simultaneous data streams from both Biopac and our systems were statistically analyzed and compared. The mean heart rate obtained by Biopac was almost the same, with maximum one beat per minute error, as the mean heart rate obtained by our system. The skin temperature delivered by Biopac corresponded to the skin temperature delivered by our system. 2.2 Software Development A schematic representation of our software prototype is shown in Fig. 6. When a biofeedback training session is started by a user, the COM Reader opens the COM port. Raw data stream comes from the OOPic microcontroller in the format: Data Separator -- Pulse Value -- Temperature Value. As a separator we used the number '2' that is not engaged in both signals. COM Reader recognizes the Data Fig. 3. Photoplethysmograph circuit used for heart rate measurement (R1=100Ω, R2= 100kΩ, R3=1.6mΩ, R4=1.6mΩ, R5= 100kΩ, R6= 1kΩ, C=2µF; Diode LED IR850, 5nm, 160 nW, λ=850nm; Phototransistor BPW42, 3mm, λp=830nm) We use the Seebeck effect for skin temperature measurement. It describes the electromotive force that exists across a junction of two dissimilar metals. We choose the Iron -- Constantan thermocouple type J with 0.25 mm diameter. It has a linear characteristic in the area around 37°C and sufficient range of values. The raw Separator and sends the next two bytes to Buffer Manager that stores pulse and temperature values. Buffer Manager also immediately sends the data to Plot Manager for display in real-time. After a time window of 10 seconds is elapsed, the content of the buffer is sent to Pulse Processor for processing. The Pulse Processor detects peaks in the arterial pulse waveform and calculates the heart rate (HR). Mean arterial pressure (MAP) is estimated from the pulse amplitude using the equation (1) which was proposed in [8]. The temperature values (T) are converted from volts values to °C according the table for thermocouple type J. HR, MAP and T are sent to Plot Manager for display. MAP a e = ⋅ b A ⋅ + ⋅ c e d A ⋅ (1) where the unknown parameters a, b, c and d can be determined by fitting correct blood pressure values to pulse amplitudes during a calibration phase. Fig. 6. Structure of a software solution for biofeedback training After a time window of 30 seconds is elapsed, the HR values corresponding to the buffer content are sent to the HRV Processor that performs a heart rate variability analysis (HRV) in the frequency band [0.015 -- 0.4] Hz by calculating the power of LF, HF components and their ratio LF/HF. The obtained HRV parameters are normalized over the total power, and then sent to Plot Manager for display. The HR, MAP, T values corresponding to the buffer content are also sent to the Model Simulator, that simulates a mathematical model of cardiovascular neural control (see section 2.3 of this paper) and estimates the level of activity of sympathetic and parasympathetic branches as well as the so-called stress score. All three values are sent to Plot Manager for display. A graphical user interface was developed in the Matlab environment according to user requirements gathered during on-site research with 30 respondents from a private relaxation center and a university (see Fig. 7). It displays measured and processed physiological data as feedback for the user. This included a percentage of stress score, with an appropriate color (red for high and green for low stress score) and a smiley icon (smiling for low, satisfied for normal and angry for high stress score). 3 Fig. 7. User interface for biofeedback training with autonomic nervous activity assessment 2.3 Mathematical Model for Biofeedback The level of activity of the autonomic nervous system can be estimated using mathematical models of cardiovascular control. The main component of such a model would be the baroreflex or baroreceptor loop, which stabilizes blood pressure by the mean of neural negative feedback with three components: an afferent or sensory component, an efferent component corresponding to autonomic nervous control and an effector component for modulating heart rate and vascular resistance [9]. We selected model equations from Ursino [10] because of their simplicity and the scientific community. We adapted the equations for our biofeedback purpose as presented below. their acceptability in The Sensory Component consists of mechanoreceptor cells located in the carotid sinus and aorta arch walls responding to pressure by stretching. Their activity is modeled using a linear derivative first-order dynamic block and a sigmoidal static block as follows. % P P n − k ab f ab = f ab,min + f ab,max ⋅ e % P P n − k ab 1 + e τ pb ⋅ % dP dt = P as + τ zb ⋅ dP as dt − % P (2) where abf is the firing rate of the carotid baroreceptors in response to the pressure sensed P% which depends on the are the upper and lower arterial pressure , f f asP ; ab,max ab,min saturation levels of firing rate with values 2.52 Hz and 47.78 Hz; τ τ are the time constants for the real pole ,pb zb 2.07 sec and 6.37 sec; and the real zero in the linear dynamic block with values nP is a model parameters, related to the value of baroreceptor pressure at the central point of the abk is a model parameter, with the dimension of pressure, related to the slope of the static function at the central point. sigmoidal function; 4 The Efferent Component consists of sympathetic and vagal nerves which project to blood vessels and heart. Their activity is modeled as exponential trends using weighted sum of afferent inputs from baroreceptors as follows. The changes in the heart rate depend of both sympathetic and vagal drive as follows. HR = ∆ HR S 60 + ∆ HR HR 0 + V f s   =   f es , ∞ + ( f es ,0 − f f k es ) ⋅ e es , ∞ es,max f ev ,0 + f ev , ∞ ⋅ e f ab − f ab ,0 k ev f ab − f ab ,0 k ev 1 + e − θ v f s f f es,max f v ev , ∞ f v = f s = f v = ( ⋅ − W f ⋅ , b s − θ s ) ab if if f f s s < ≥ f f es,max es,max ∆ d HR S dt ∆ d HR V dt = = 1 τ HR S 1 τ HR V ( ⋅ −∆ HR S + σ HR S ) ( ⋅ −∆ HR V + σ HR V ) (5) (3) σ HR S σ HR V  G =   G = HR v ⋅ ln f [ s ( t D − HR S HR S ) − f es,min + 1] if f s if f s ≥ < f f es,min es,min 0 ) ⋅ f v ( t D − HR V 1 HR 0 = 1.97 9.5 10 − ⋅ − 3 ⋅ age where sf is the firing rate of efferent sympathetic fibers to vessels and heart; vf is the firing rate of efferent vagal es , , , f f k k es,max fibers to heart; ∞ are constants with values 0.06, 60 Hz, 2.1 Hz, 7 Hz, 2.1 Hz and 3.15 Hz; vθ θ are offset ,b spW is a synaptic weight with value 0.3; terms for sympathetic and vagal neural activation with ,s f , ∞ ev , , ev es values -49 Hz and -0.68 Hz. f es ,0 , f ev ,0 are model to the related parameters, rate of sympathetic and vagal nerves respectively at the absence of sf is the so-called sympathetic tone, i.e. a quantification of sympathetic activity in the normalized any sensory input. intrinsic firing range [0 .. 1]. Similarly, vf is the parasympathetic tone. The Effectors Component consists of the heart sinoatrial node and peripheral blood vessels. The changes in skin temperature as response to sympathetic activity are modeled to include a pure latency, a monotonic logarithmic static function, and low-pass first order dynamics. ( ) T t s ∆ d T s dt = ∆ ( ) T t s + T s ,0 = 1 τ T S ( ⋅ −∆ + T s σ T s ) (4) σ T s  =   G ln f [ ⋅ T S − ( t D T S s ) − f es,min + 1] 0 if f if f s s ≥ < f f es,min es,min where sT is the skin temperature in ºC, related to the value as model parameter; resistance of peripheral vessels with ,0sT being its baseline STτ is a time constant with value STD is the time delay for sympathetic response to take effect with STG is a constant gain factor with value 1.94; 6; value 5 sec; esf , min is a threshold for sympathetic stimulation with value 2.66 Hz. where HR is the heart rate in beats/min with intrinsic duration of 1 cardiac cycle, depending on the age ; HR being the 0 τ τ are a time constants with values 2 and 1.5; , HR S HR V G HR S , G HR V are constant gain factors with values -0.13 and 0.09; D , D HR V HR S are the time delays for sympathetic and parasympathetic responses to take effect with values 2 sec and 0.2 sec. 2.4 Experiments Three healthy men and three healthy women (age 37±13 years) with different employments and stress environments were divided into two randomly selected groups. The first one undergoes whole biofeedback training and the second one was a non-training control group. The control and trained group were managed according to training plans described in the Fig. 8 and 9. All subjects were instrumented with the hardware device we have developed temperature measurement (see section 2.1). for arterial pulse and skin The goal of the training for both groups was to improve the ability to calm down and relax. The methods for relaxation were deep diaphragmatic breathing and enhanced peripheral capillary return. Deep diaphragmatic breathing was performed at a rate of 6 cycles per minute in a comfortable sitting position with one hand placed on the upper chest and the other just below your rib cage in order to feel the diaphragm moving. The subjects were instructed to tighten the stomach muscles, letting them fall inward as they exhale through pursed lips, while the hand on the upper chest remains as still as possible. Enhanced peripheral capillary return was upheld by repeating autogenic phrases such as "I have warm stones in my hands which heat them". Relaxation was further enhanced in the training group by gentle massage and few minutes resting in yoga asana Savasana which has the same sedative effect as some kind of antidepressants. The control group did not undergo massage and yoga. 5 The principle of training was based on scoring the overall stress state in order to provide the learner with combined information about his or her heart rate, skin temperature, heart rate variability, level of sympathetic or parasympathetic activity and stress score in real time. The control group did not have access to that combined information. Instead control subjects had to achieve the biofeedback goal without being presented with any information about their physiological state. Each biofeedback session had 4 phases. The pure measurement phase is for the subject to get familiarized with the experimental setup. The deep diaphragmatic breathing phase is preparation phase where subjects get relaxed by activating their parasympathetic system via deep breathing and fictive image of warm hands. The resting phase is a pause prior to the final phase. The Self-calming phase is the active biofeedback session, where the subject consciously tries to take control of its physiological variables while reducing his stress score, sympathetic tone and low frequency components in heart rate variability, and increasing his parasympathetic tone and high frequency components in heart rate variability at the same time. Fig. 8. Experimental schedule of the training group. Fig. 9. Experimental schedule of the control group. 2.5 Parameters estimation Beat-to-beat heart rate ( HR meas ) and skin temperature ( measT ) were directly extracted from the pulse plethysmograph signal which was recorded during the all four phases of the biofeedback training. The obtained HR time series was analyzed in the frequency domain using Fourier transformation in order to calculate the normalized power of low and high frequency components of heart rate variability (LF, HF). Mean arterial pressure ( MAP meas ) was estimated from the amplitude of peaks in the pulse plethysmograph signal using the mathematical equation (1). In order to compute the sympathetic tone sf and the parasympathetic tone vf , the mathematical model described in section 2.3 was implemented in the Simulink software environment and model parameters were estimated by formulating and resolving a least-squares the errors between optimization problem minimizing measured skin temperature heart rate, and HR smea , MAP , T meas meas and simulated heart rate, and skin temperature HR P T . The corresponding cost function, , ,as s i.e. the discrepancy between measured and simulated values is defined as optimization criteria as follows, where n is the number of data points. J = n ∑ [ i=1 HR meas ] (i) - HR(i) 2 + [ MAP meas ] (i) - P (i) 2 as + [ T meas ] (i) - T (i) 2 s (5) The cost function was minimized in the Matlab software environment with parameters values shown in table 1 for a representative subject. Model Parameter Value nP abk ,0esf ,0evf ,0sT SHRG VHRG 92 mmHg 11.7 mmHg 16.1 Hz 3.14 Hz 25 ºC -0.13 0.09 Tab. 1. Typical values of model parameters The simulated sympathetic and parasympathetic tones were combined with variables extracted from experimental data in order to calculate the stress score using the formula (6) below. Sympathetic tone, high fluctuations in heart rate, average skin temperature close to 37 ºC and average heart rate less than 70% of the intrinsic heart rate contribute to lower stress score, meaning that the subject is fully relaxed under predominant activity of the parasympathetic branch of the autonomic nervous system, compare to minimal activity of the sympathetic branch. 6 SC = 100 ⋅     − 1     f HF f v + − 3. Results HR 0 − 60 HR meas 0.7 ⋅ HR 0 + T ,0 s 37 − LF + s 3 (6)          In order to evaluate two consecutive phases of the effectiveness of our biofeedback method which is based on autonomic nervous assessment, we compared the change in different variables between the biofeedback training for the control group, the trained group before relaxation and the trained group after relaxation. Especially we are interested in the capability of the trained group to more effectively reduce its stress score and increase its parasympathetic tone during self calming (phase 4). Heart rate decreased during deep diaphragmatic breathing (phase 2) for all groups. This trend was also observed in the trained group during self calming. The control group displayed an increasing heart rate instead, as shown in Fig. 10. Fig. 10. Change in the average heart rate. in The changes the normalized power of low frequency components of heart rate variability, which reflect sympathetic activity, are the most expressive for the trained group. Subjects in the trained group had visual information about their autonomic nervous activity and could reduce sympathetic activity in phase 4 (see Fig. 11) while enhancing vagal activity. In Fig. 12, the power of high frequency components of heart rate variability, which are known to reflect vagal activity and a state of relaxation, significantly increases. The control group did not exhibit any change. Fig. 12. Changes in the normalized power of high frequency components (HF) of heart rate variability. As shown in Fig. 13, the control group had the highest stress score. Stress decreased during biofeedback training with our software solution. Relaxation prior to biofeedback training provided additional reduction of stress score. Fig. 13. Stress score achieved during self calming. Differences between the control and trained group are obvious as explained in previous paragraphs. The trained groups show better results (reduced stress score, reduced sympathetic tone, increased parasympathetic activity). It was also observed that passive relaxation techniques are valuable for effective biofeedback training. We believe that our complete and functional solution for biofeedback training can be used in alternative medicine, i.e. therapy with limited adverse effects. Future work could be in the design of a better photosensor gripping on the finger in order to reduce noise in the pulse signal. Moreover we should improve the mathematical model of autonomic nervous control and include additional neural pathways that are involved in sympathetic and parasympathetic function. We also plan to test the solution on more volunteers in order to bring a solid statistical proof of its significant effectiveness in biofeedback training. A further room of improvement would be the inclusion of environmental parameters to our solution such as ambient temperature, humidity and atmospheric pressure which can be very influential on the autonomic nervous system of human subjects. Fig. 11. Changes in the normalized power of low frequency components of heart rate variability. References [1] Schwartz, M and Andrasik, F., Biofeedback: A Practitioner's Guide. 3rd Edition. s.l. : The Guilford Press, 2005. ISBN 1593852339. 7 [2] Arns, M, et al., "Efficacy of neurofeedback treatment in ADHD: the effects on inattention, impulsivity and hyperactivity: a meta-analysis." Clin EEG Neurosci., 2009, Issue 3, Vol. 40, pp. 180-9. [3] Unglaub, D., Human Physiology: An Integrated [ed.] 4th Edition. s.l. : Pearson, 2009. Approach. 0805368493. [4] Malik, M., "Heart Rate Variability. Standards of Measurement, Physiological Interpretation and Clinical Use." Circulation, 1996, Issue 5, Vol. 93, pp. 1043-1065. [5] Goldberger, J J, et al., "Dissociation of heart rate variability from parasympathetic tone." Am J Physiol Heart Circ Physiol, 1994, Issue 5, Vol. 266, pp. H2152-H2157. [6] Batzel, J, et al., Cardiovascular and Respiratory Systems -- Modeling, Analysis and Control. Philadelphia : Society for Industrial and Applied Mathematics, 2007. pp. 6-16. ISBN 0-89871-617-9. [7] Allen, its application in clinical physiological measurement." Physiol Meas, 2007, Issue 3, Vol. 28, pp. R1-39. J., "Photoplethysmography and [8] Shaltis, P, Reisner, A and Asada, H., "Calibration of the photoplethysmogram to arterial blood pressure: capabilities and for continuous pressure monitoring." 2005. Conf Proc IEEE Eng Med Biol Soc. Vol. 4, pp. 3970-3. limitations [9] Ottesen, J T., "Modelling of the baroreflex- Journal of feedback mechanism with Mathematical Biology, 1997, Issue 1, Vol. 36, pp. 41-63. time-delay." [10] Ursino, M., "Interaction between carotid baxoregulation and the pulsating heart: a mathematical model." Am J Physiol Heart Circ Physiol, 1998, Vol. 275, pp. 1733-1747.
1602.01889
2
1602
2016-06-11T01:45:12
Discovering Neuronal Cell Types and Their Gene Expression Profiles Using a Spatial Point Process Mixture Model
[ "q-bio.NC", "stat.ML" ]
Cataloging the neuronal cell types that comprise circuitry of individual brain regions is a major goal of modern neuroscience and the BRAIN initiative. Single-cell RNA sequencing can now be used to measure the gene expression profiles of individual neurons and to categorize neurons based on their gene expression profiles. While the single-cell techniques are extremely powerful and hold great promise, they are currently still labor intensive, have a high cost per cell, and, most importantly, do not provide information on spatial distribution of cell types in specific regions of the brain. We propose a complementary approach that uses computational methods to infer the cell types and their gene expression profiles through analysis of brain-wide single-cell resolution in situ hybridization (ISH) imagery contained in the Allen Brain Atlas (ABA). We measure the spatial distribution of neurons labeled in the ISH image for each gene and model it as a spatial point process mixture, whose mixture weights are given by the cell types which express that gene. By fitting a point process mixture model jointly to the ISH images, we infer both the spatial point process distribution for each cell type and their gene expression profile. We validate our predictions of cell type-specific gene expression profiles using single cell RNA sequencing data, recently published for the mouse somatosensory cortex. Jointly with the gene expression profiles, cell features such as cell size, orientation, intensity and local density level are inferred per cell type.
q-bio.NC
q-bio
Discovering Neuronal Cell Types and Their Gene Expression Profiles Using a Spatial Point Process Mixture Model Furong Huang Animashree Anandkumar UC Irvine UC Irvine Christian Borgs Microsoft Research [email protected] [email protected] [email protected] Jennifer Chayes Ernest Fraenkel Michael Hawrylycz Microsoft Research MIT Allen Institute Ed Lein Allen Institute [email protected] [email protected] [email protected] [email protected] Alessandro Ingrosso Srinivas Turaga Politecnico di Torino HHMI Janelia Research Campus [email protected] [email protected] Abstract Cataloging the neuronal cell types that comprise circuitry of individual brain regions is a major goal of modern neuroscience and the BRAIN initiative. Single-cell RNA sequencing can now be used to measure the gene expression profiles of individual neurons and to categorize neurons based on their gene expression profiles. While the single-cell techniques are extremely powerful and hold great promise, they are currently still labor intensive, have a high cost per cell, and, most importantly, do not provide information on spatial distribution of cell types in specific regions of the brain. We propose a complementary approach that uses computational methods to infer the cell types and their gene expression profiles through analysis of brain-wide single-cell resolution in situ hybridization (ISH) imagery contained in the Allen Brain Atlas (ABA). We measure the spatial distribution of neurons labeled in the ISH image for each gene and model it as a spatial point process mixture, whose mixture weights are given by the cell types which express that gene. By fitting a point process mixture model jointly to the ISH images, we infer both the spatial point process distribution for each cell type and their gene expression profile. We validate our predictions of cell type-specific gene expression profiles using single cell RNA sequencing data, recently published for the mouse somatosensory cortex. Jointly with the gene expression profiles, cell features such as cell size, orientation, intensity and local density level are inferred per cell type. 1 Introduction 1.1 Motivations and Goals The human brain comprises about one hundred billion neurons and one trillion supporting glial cells. These cells are specialized into a surprising diversity of cell types. The retina alone boasts well over 50 cell types, and it is an active area of research to perform a census of the various neuronal cell types that comprise the central nervous system. Many criteria have been used to categorize neuronal cell types, from neuronal morphology and connectivity to their functional response properties. Neurons can also be categorized based on the proteins they make. Immunohistochem- istry has been used with great success for many decades to differentiate excitatory neurons from inhibitory neurons by labeling for known proteins involved in the synthesis and regulation of glutamate and GABA, the primary excitatory and inhibitory neurotransmitters respectively. More recently, there has been an effort to systematically measure the complete transcriptome of single neurons. Single-cell RNA sequencing (RNA-Seq) is an extremely powerful technique that can quantitatively determine the ex- pression level of every gene that is expressed in individual neurons. This so-called transcriptome or gene expression 1 / transcription profile can then be used to define cell types by clustering. A recent study produced the most compre- hensive census of cell types to date in the mouse somatosensory cortex and hippocampus by performing single-cell RNA-Seq on over 3000 neurons [12]. While this study is quite exciting, tyring to replicate it for all brain regions might well require the equivalent of a thousand such experiments. Thus, it is likely that the unprecedented insights that RNA-Seq can provide will be slow to arrive. More importantly, single cell sequencing methods are not currently able to capture the precise three-dimensional location of the individual neurons. Here we propose a complementary approach that uses computational strategies to identify cell types and their spatial distribution by re-analysing data published by the Allen Institute for Brain Research. The Allen Brain Atlas (ABA) contains cellular resolution brain-wide in-situ hybridization (ISH) images for 20,000 genes1. ISH is a histo- logical technique that labels the mRNA in all cells expressing the corresponding gene in a manner roughly proportion to the gene expression level. An example of an ISH image can be seen in figure 1(a). The ABA contains genome-wide and brain-wide ISH images of the adult mouse brain. These images were gener- ated by slicing the brain into a series of 25 µm thin sections and performing ISH. Image series of ISH performed for different genes come from different mouse brains, since ISH can only be performed for one gene at a time. The ISH image series for different genes were then computational aligned into a common reference brain coordinate system. Such data have been productively used to infer the average transcriptomes corresponding to different brain regions. It is commonly thought that the ABA cannot be used to infer the transcriptomes of individual cells in a given brain region since mouse brains cannot be aligned to the precision of a single cell. This is because there is individual variation in the precise number and location of neurons from brain to brain. However, we expect that the average number and spatial distribution of neurons from each cell type to be conserved from brain to brain, for a given brain area. More concretely, we might expect that parvalbumin-expressing (PV) inhibitory interneurons in layer 2/3 of the mouse somatosensory cortex comprise approximately 7% of all neurons and have a conserved spatial and size distribution from brain to brain. We use this fact to derive a method for simultaneously inferring the cell types in a given brain region and their gene expression profiles from the ABA. We propose to model the spatial distribution of neurons in a brain as being generated by sampling from an unknown but consistent brain-region and cell-type dependent spatial point process distribution. And since each gene might only be expressed in a subset of cell types, an ISH image for a single gene can be thought of as a mixture of spatial point processes where the mixture weights represent the individual cell types expressing that gene. We infer cell types, their gene expression profiles and their spatial distribution by unmixing the spatial point processes corresponding to the ISH images for 1743 genes. This is in notable contrast to the information provided by single-cell RNA sequencing which can only measure the gene expression profile of individual cells to high accuracy but where, due to the destructive measurement process, all information about the spatial position and distribution of cell types is lost. 1.2 Previous Work Allen Brain Atlas (ABA) [9] is a landmark study which mapped the gene expression of about 20,000 genes across the entire mouse brain. The ABA dataset consists of cellular high-resolution 2d imagery of in-situ hybridized series of brain sections, digitally aligned to a common reference atlas. However, since the in-situ images for each gene come from different mouse brains and since there is significant variability in the individual locations of labeled cells, it is not possible to register brain-wide gene expression at a resolution higher than about 250µm. Therefore, the cellular resolution detail was down-sampled to construct a coarser 3d representation of the average gene expression level in 250µm × 250µm × 250µm voxels. The coarse-resolution averaged gene expression representation has been widely used and analyzed to understand differences in gene expression at the level of brain region. Hawrylycz et al [5] analyzed the correlational structure of gene expression at this scale, across the entire mouse brain. However, due to the poor resolution of the average gene expression representation, it has proven challenging to use the ABA to discover the microstructure of gene expression within a brain region. To address this issue from a complementary perspective, Grange et al [3] used the gene expression profiles of 64 known cell-types, combined with linear unmixing to determine the spatial distribution of 1 Although the Atlas contains ISH data for approximately 20,000 distinct mouse genes, we focus on the top 1743 reliable genes whose sagittal and coronal experiments are highly correlated. 2 these known cell-types. However, such an approach can be confounded by the presence of cell-types whose expression profiles have yet to be characterized, and limited by the resolution of the averaged gene expression representation. In contrast to previous approaches, we aim to solve the difficult problem of automatically discovering the gene expression profiles of cell-types within a brain region by analyzing the original cellular resolution ISH imagery. We propose to use the spatial distributions of labeled cells, and their shapes and sizes, which are a far richer representation than simply the average expression level in 250µm × 250µm × 250µm voxels. This spatial point process is then un-mixed to determine the gene expression profile of cell types. Most previous work on unmixing point process mixtures adopted parametric generative models where the point process is limited to some distribution family such as Poisson or Gaussian [6, 7]. However, since we are not interested in building a generative model of a point process, but rather care more about inferring the mixing proportions (gene expression profile), we take a simpler parameter-free approach. This approach models only the statistics of the point process, but is not a generative model, and so cannot be use to model individual points/cells. 2 Modeling the Spatial Distribution of Cell-types Using Spatial Point Pro- cess Features Most analyses of the ABA in situ hybridization dataset have utilized a simple measure of average expression level in relatively large 250µm × 250µm × 250µm voxels of brain tissue. Due to the large volume over which the expression level is averaged, such a representation cannot distinguish between large numbers of cells expressing small amounts of RNA vs. small numbers of cells expressing large amounts of RNA. All information about the spatial organization of labeled cells, their shapes, sizes and spatial density are lost and summarized by a single scalar number. Here, we describe a more sophisticated representation of the labeled cells in an ISH image based on marked spatial point processes. 2.1 The Marked Spatial Point Process Representation of ISH Images Our approach requires processing the high-resolution ISH images to detect individual labeled cells and their visual characteristics. We developed a cell detection algorithm described in the Supplementary section. Our algorithm additionally also estimates the expression level of each detected cell, its shape, size and orientation. Figure 1(a) and Figure 1(b) illustrate the results of our cell detection algorithm. Since cell-types differ not only in terms of gene expression pattern, but also display a diversity of shapes, sizes and spatial densities, we sought to characterize these properties. We measured: (1) cell size s = [r1, r2]: the radius in two principal directions of an ellipse fit to each cell; (2) cell orientation o: the orientation of the first principle axis of the ellipse; (3) gene intensity level p: intensity of labeling of a cell relative to the image background; (4) spatial distribution c: the number of cells within a local area centered around the cell, which can be regarded as a measure of the local cell density. The collection of detected cells within an atlas-defined brain region, along with their features, constitutes a marked spatial point process. This point process is considered "marked", because each point is characterized by the shape, size, expression level and local density features, in addition to just their location in space. 2.2 A Model-free Approach to Representing Spatial Point Processes Using Joint Feature Histograms The statistical modeling of repulsive spatial point processes such as those that arise in biology is non-trivial, and many generative models such as determinantal point processes [8]and Matern point processes have high computational complexity. But since we are not interested in directly modeling the individual labeled cells, but instead in modeling only their aggregate spatial statistics, and in inferring their gene expression profiles, we can take a simpler approach. We use a joint histogram simple statistics of the collection of detected cells to characterize the underlying point process from which they are drawn. This is an empirical moment approach which side-steps the need to carefully define a generative point process distribution. 3 Extract Point Process: 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 PSfrag replacements 0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 (a) Patch from gene Pvalb slice (b) Cell detection and extraction of spatial point process features Joint Histogram: Pvalb Rasgrf2 PSfrag replacements cell number PSfrag replacements cell number (c1) Size π/6 2π/6 3π/6 4π/6 5π/6 (c2) Orientation PSfrag replacements cell number PSfrag replacements cell number (c3) Expression level (c4) Cell counts in 100 µm radius Discover Cell Types: (d) Point process histogram representation: [xm n ] ∈ R+ NG×NF (f) LDA model for inferring cell types Figure 1: Overview of the proposed framework - Discovering Neuronal cell Types via Un-mixing of Spatial Point Process Mixtures. (a) & (b) An in situ hybridization image for gene Pvalb along with detected cells. (c) Marginalized point process feature histograms for genes Pvalb and Rasgrf2. Note that size denotes the principal axis diameter. We have NG genes and 4d joint histogram with NF bins. As we describe in the next section, we propose to model the point process measured from the ISH image for each gene as a mixture of point processes belonging to individual cell-types. For this, we use a linear mixing model, the Latent Dirichlet Allocation model. The use of this model is greatly simplified if we carefully choose our feature representation such that the linear mixture of point processes results in a linear mixture of histogram statistics. This is clearly the case for the features we have chosen. For instance, if we sample equally from two point process distributions P1 and P2 with average densities of d1 and d2, the addition of these two point processes P = P1 + P2 results in the addition of the two densities d = d1 + d2. This is not the case for second order features, such as the distances to the nearest neighbors, which would have a more nonlinear relationship. In figure 1(c), we display marginal histograms corresponding to the joint histogram for two genes, Pvalb and 4 Rasgrf2, which are well-known markers for a specific class of inhibitory and excitatory cortical neuronal cell-types respectively. 3 Un-mixing Spatial Point Processes to Discover Cell-types 3.1 Generative Model: A Variation of Latent Dirichlet Allocation The spatial point process histogram representation of the ABA ISH dataset results, for each brain region, is an NF ×NG matrix [xm n ], where NF is the total number of histogram bins (henceforward called the number of histogram features) 2, NG is the number of genes, and xm n is the number of cells expressing gene n in histogram bin m. We model the gene-spatial histogram matrix [xm n ] by assuming it is generated by a Variation of Latent Dirichlet Allocation (vLDA) [2] model of cell types. This matrix factorization based latent variable model assumes that the ISH histograms are generated from a small number of cell-types, K, and each cell-type i is associated with a type- dependent spatial point process histogram hi and a gene expression profile βi. n xm spatial distribution) is as follows: Let Lm =PNG Our generative model for each histogram bin m (characterizing a particular bin in the size/ orientation/ gene profile/ n be the detected number of cells in the joint histogram bin m. For each cell l in this bin, its cell-type t is sampled from the multinomial distribution hm. And given the cell-type t of cell l, the genes n expressed by this cell are sampled from a multinomial distribution given by the type-dependent gene expression profile/distribution βt. For a given gene n and histogram bin m, this generative process determines n . the number of cells that would be detected xm We further place a Dirichlet prior over hm ∼ Dir(α), with the concentration parameter α which determines the prior probability over the number of cell-types present in a given histogram bin m. This prior represents our prior knowledge of how many cell-types express each gene, and also how well our feature representation separates cells of different types into different histogram bins. In principle, we could generalize this to be a gene-specific prior, if we had such information available. We could also use α to incorporate information about our prior knowledge over the distribution of cells from each cell-type, for instance that excitatory neurons greatly outnumber inhibitory neurons in a roughly 5 : 1 ratio. We now describe how we estimate the model parameters -- the cell-type specific multinomial gene expression profile β and the cell-type specific spatial point process histogram h from the gene-specific spatial point process histograms measured from the ISH images. 3.2 Estimating the Cell-type Dependent Gene Expression Profile β After testing several estimation methods for the parameters of our model, we found that non-negative matrix factor- ization (NMF) performed well in estimating the cell-type specific gene expression profiles β, see Figure 2a. We solve the following optimization problem: min β,h NFXm NGXn (xm n − KXt hm t βt nLm)2, s.t. βt n ≥ 0, βt n = 1, hm t ≥ 0, NGXn KXt hm t = 1 (1) Here, the non-negativity and sum-to-one constraints on hm n ensure that h and β result in properly normalized multinomial distributions. While this estimation procedure results in joint estimates for h and β, it does not enforce the Dirichlet prior over h. So we refine our NMF-derived estimates for h using variational inference [2]. t and βt 2Note that there are two types of features -- the features characterizing each detected cell, and the features characterizing the collection of detected cells that constitute a single sample from a spatial point process 5 3.3 Estimating the Cell-type Dependent Spatial Point Process Histogram h We use a standard maximum likelihood estimation procedure for h [2]. Iteratively, we refine the inference of the cell type membership hm ∈ ∆k under each joint histogram feature m. We update hm i until convergence [11]. hm i ← 1 Lm +PK t αt xm n NGXn=1 KPl=1 hm i βi n hm l βl n + αi, ∀i ∈ [K], m ∈ [NF ] (2) Recall that the Dirichlet prior α encodes the number of cell-types that we expect on average to express each gene. We set α to be a symmetric Dirichlet with α1 = α2 = . . . = αK, andPt αt = 0.01 for all cell-types t. In practice, we observe that our estimates of h are fairly insensitive to the specific choice for α as long asPt αt is small enough. The smaller α is, the fewer cell-types expressing a given gene we expect to observe in a single histogram bin. 4 Results and Evaluation 4.1 Implementation Details We tested our proposed cell-type discovery algorithm using the high-resolution in situ hybridization image series for 1743 of the most reliably imaged and annotated genes in the ABA. Individual cells were detected in the cellular resolution ISH images using custom algorithms (detailed in Supplementary Information). For each detected cell, we fit ellipses and extract several local features: (a) size and shape represented as the diameters along the principle axes of the ellipse, (b) orientation of the first principle axis, (c) gene intensity level as measured by the intensity of labeling of the cell body, and (d) the number of cells detected with-in a 100 µm radius around the cell, which is a measure of the local cell density. We aligned the ISH images to the ABA reference atlas and, for this paper, focused our attention on cells in the somatosensory cortex, since independent RNA-Seq data exist for this region the can be used to evaluate our approach. We computed joint histograms for the collection of cells found with-in the somatosensory cortex, resulting in a spatial point process feature vector of NF = 10010 histogram bins per gene. Synthetic experiment: The vLDA model we proposed is then fit to NG × NF gene point process histogram matrix to estimate the cell-type gene expression profile matrix β using the non-negative matrix factorization (NNMF) algorithm. The reason why we choose NNMF over Variational Inference (which is a popular approach for LDA) for β estimation is that NNMF produces more accurate β estimation in simulated data, illustrated in Fig 2a. In the synthetic experiment, we simulate point process data ( with some predefined golden standard β) and use the data to estimate bβ. The errors were computed after pairing the estimated columns of β with a closest golden standard β column via hypothesis testing. Note that the columns of β are normalized to 1, so the errors are bounded. 4.2 Evaluating Cell-type Gene Expression Profile Predictions A recent study performed single-cell RNA sequencing on 1691 neurons isolated from mouse somatosensory cortex. We use this dataset to evaluate the quality of the cell-types we discover. The single cell RNA-seq data, G := [g1g2 . . . gNC ] ∈ RNG×NC , contains the gene expression profiles for NC = 1691 cells. We infer the cell types hi for these cells using equation (2), and then compute the likelihood Li of observing each for each cell under our estimated cell-type dependent gene expression profile matrix β using equation (4). We can then evaluate the perplexity, a commonly used measure of goodness of fit under the vLDA model, of single cell RNA-seq data on the model we learned from our spatial point process data. The perplexity score is a standard metric, which is defined as the geometric mean per-cell likelihood. It is a monotonically decreasing function of the log-likelihood L(G) of test data G. perplexity(G) = exp(−PNC PNC i=1 Li i=1 log p(gi) 6 ) (3) PSfrag replacements 0.2 Permute β VI estimated β NNMF estimated β NNMF robust k 2.5 Spatial point process (ours) ( permuted ) Average expression level (baseline) ( permuted ) PSfrag replacements 2.0 e p y T r e P r o r r E 0.15 0.1 0.05 0 3 10 20 30 40 50 Number of Cell Types Spatial point process (ours) Average expression level (baseline) Spatial point process (ours, permuted) Average expression level (baseline, permuted) 60 80 70 90 100 Perplexity Score 1.5 1.0 0.5 (a) Validate NNMF Method Number of Cell Types (b) Validate Point Process Data Figure 2: (a) Synthetic Experiment : comparison of Non-negative Matrix Factorization (NNMF) with Variational Inference (VI) on simulated point process cell data using known gene expression profile β. An additional robustness test of NNMF is done to see how good the algorithm is when a wrong number of cell types K is input. A permutation test (shuffling the gene expression levels between cell) is done to access statistical significance. Comparing with permute test shows that our cell-types are significantly different from chance. Error per type is computed by pairing the columns of estimatedbβ with the columns of the ground-truth β. (b) Comparison of gene expression profiles recovered for cell-types in the somatosensory cortex by fitting an LDA model using spatial point process features (ours) vs the standard average gene expression level feature (baseline). Our features provide a significantly better match, with lower perplexity, to ground truth single-cell RNA sequencing derived transcriptomes. A permutation test is done to access statistical significance. Perplexity is computed by matching to surrogate single-cell RNA transcriptomes by shuffling the gene expression levels between cells. Comparing with permute test shows that our cell-types are significantly different from chance. where the likelihood is evaluated as p(gmhm, α, β) = Γ (Pi αi) Qi Γ (αi) (hm i )αi−1 kYi=1 LmYj=1 kXi=1 NGXn=1 δgm j ,en hm i βi n! . (4) where δi,j is the Kronecker delta, δi,j = 1 when i = j and 0 otherwise. en is the nth basis vector. 4.3 Comparison to Standard Average Gene Expression Features Baseline and a Permuta- tion Test for Significance Here we demonstrate the superiority of our method and its statistical significance in two ways. First we compared the perplexity of the single-cell RNA seq dataset G under our model (figure 2b, solid blue) against the perplexity of a surrogate dataset with the same marginal statistics, but whose gene-cell correlations were destroyed (figure 2b, dashed blue). We generated this surrogate dataset by randomly permuting the gene expression levels for each gene across cells. This permuted dataset had a significantly higher (worse) perplexity than the true single-cell dataset. This demonstrates that our model trained to un-mix the ISH-derived spatial point processes discovered cell-types whose gene expression profiles are significantly better match to single-cells than by chance. We also compared the predictions of cell-type gene expression profiles derived by un-mixing our spatial point process features against gene expression profiles derived by un-mixing the more standard 250µm × 250µm × 250µm averaged gene expression level features. We see a very large improvement in perplexity by switching from the standard simple averaging of gene expression, to extracting spatial point process features (figure 2b). The single-cell RNA seq dataset analysis from figure 2b shows that the perplexity of our recovered cell-types rapidly flattens after we recover approximately 10 clusters (K = 10). 7 [2,4] [3,3] [3,4] [3,5] [3,6] [3,7] [4,4] [4,5] [4,6] [4,7] [5,7] [6,6] [6,7] [7,7] [7,8] [7,9] [8,8] [8,9] [9,9] Gad1 Sp8 Tox3 Nkx2-1 Lhx6 Pax6 Dlx5 Arx Dlx2 Dlx1 Elavl2 Sp9 Tbr1 Foxp2 Tshz2 CA1Pyramidal Stat4 Ascl1 Cux2 Neurod1 Mef2c Oligodendrocytes S1Pyramidal Ependymal Microglia Astrocytes Interneurons Endothelial M ural PSfrag replacements PSfrag replacements Ptrf Cldn5 Maf Hcls1 Spi1 Myb Fhl1 Aldoc Sall3 Sox21 Mbp CA1 Pyramidal Etv6 Sox10 St18 Olig2 Oligodendrocytes S1 Paramidal Ependymal Microglia Astrocytes Interneurons Endothelial M ural Figure 3: Estimated memberships β on marker genes for 8 cell types. These marker genes are used to label the columns of the membership matrix. PSfrag replacements CA1Pyramidal Cell Fraction Cell Fraction 4.4 A Brief Analysis of Recovered Cell Types in Somatosensory Cortex Cell Fraction Cell Fraction π 6 In this section we describe the representative spatial point process statistics and gene expressions for 8 cell-types we recovered. We attempted to align our 8 clusters to cell-types defined by [12] in the single-cell RNA sequencing paper. We found high overlap in the gene expression profiles for all 8 clusters with known cell-types defined in [12], Interneurons, S1 Pyramidal, Mural, Endothelial, Microglia, Ependymal, Astrocytes and Oligodendrocytes, in Figure 3. 4π 6 3π 6 2π 6 5π 6 Axis 1 Axis 2 CA1Pyramidal 150 100 50 9 8 7 6 5 4 3 0.9 0.8 0.7 0.6 0.5 0.4 0.3 CA1Pyramidal 50 40 30 20 10 CA1Pyramidal 2 n s I n ter n e u r o y ra P o O li g 1 S d r o c y te s d ial m M icr o d e n E g lia o t h elial A str o c y te s p e n d n E al m M y d u ral 0 n s I n ter n e u r o y ra P o O li g 1 S d r o c y te s d ial m M icr o d e n E o t h elial g lia A str o c y te s p e n d n E y d m al M u ral 0.2 n s I n ter n e u r o y ra P o O li g 1 S d r o c y te s d ial m M icr o d e n E o t h elial g lia A str o c y te s p e n d n E al m M y d u ral (a) Cell diameter in principal axes (b) Orientation (c) Intensity 0 S 1 al d r o c y te s d ial m M icr o d e n E g lia o t h elial A str o c y te s p e n d n n s I n ter n e u r o y ra P o O li g (d) Cells in 100 µm radius u ral M m y d E Figure 4: Figure of 5% and 95% percentile estimated cell features for 8 cell types we detected. Inference is performed on the Spatial point process histograms data we estimated. The estimate of β was combined with MLE to infer the cell-type specific spatial point process representation hm . l In examining the spatial point process distributions that we predict for each of these cell types, we discover that while the distribution of cell body orientations is quite broad and similar across cell types, the cell count distribution, which is a measure of cell density, varies in a systematic way from one cell type to another. Fig 4d shows that inhibitory Interneurons are less dense than S1Pyramidal neurons. This is consistent with their known prevalence, roughly 20% of all neurons are GABAergic interneurons [10], while the remaining 80% are excitatory glutamatergic pyramidal neurons. As expected, this excitatory neuronal category of S1Pyramidal is the most common and hence most dense class of neuronal cells. They also have slightly larger cell bodies, compared to interneurons, as can be seen in Fig 4a. The remaining 6 cell types correspond to various glial sub-types. 8 5 Conclusion We developed a computational method for discovering cell types in a brain region by analyzing the high-resolution in situ hybridization image series from the Allen Brain Atlas. Under the assumption that cell types have unique spatial distributions and gene expression profiles, we used a varied latent Dirichlet allocation (vLDA) based on spatial point process process mixture model to simultaneously infer the cell feature spatial distribution and gene expression profiles of cell types. By comparing our gene expression profile predictions to a single-cell RNA sequencing dataset, we demonstrated that our model improves significantly on state of the art. The accuracy of our method relies heavily on the assumption that cell-types differ in their spatial distribution, and that our point process features perform a good job of distinguishing these differences. Thus the performance of our method can be improved by better estimates of better features. We would expect our method to perform better for large brain areas, which can be more accurately aligned, and which have more cells to estimate point process features. There are several modifications to our vLDA model which might improve the faithfulness of our generative model to the biology. We place a symmetric Dirichlet prior over cell-type multinomial distribution hm for a given histogram bin m. This assumes that the number of cell-types expressing each gene is the same for all genes. But since some genes are expressed more commonly and non-specifically than others, we might expect a gene-specific prior to be a better model. Further, the symmetric Dirichlet assumes that all cell-types have equal proportions of cells. But evidence suggests that excitatory neurons are more common than inhibitory neurons in cortex [4], and using a non-uniform Dirichlet prior could account for this. References [1] Website: 2014 allen institute for brain science. allen mouse brain atlas [internet]. Available from: http://mouse.brain-map.org/. Accessed: 2014-11-06. [2] David M Blei, Andrew Y Ng, and Michael I Jordan. Latent dirichlet allocation. the Journal of machine Learning research, 3:993 -- 1022, 2003. [3] Pascal Grange, Jason W Bohland, Benjamin W Okaty, Ken Sugino, Hemant Bokil, Sacha B Nelson, Lydia Ng, Michael Hawrylycz, and Partha P Mitra. Cell-type -- based model explaining coexpression patterns of genes in the brain. Proceedings of the National Academy of Sciences, 111(14):5397 -- 5402, 2014. [4] Kenneth D Harris and Thomas D Mrsic-Flogel. Cortical connectivity and sensory coding. Nature, 503(7474):51 -- 58, 2013. [5] Mike Hawrylycz, Lydia Ng, Damon Page, John Morris, Chris Lau, Sky Faber, Vance Faber, Susan Sunkin, Vilas Menon, Ed Lein, et al. Multi-scale correlation structure of gene expression in the brain. Neural Networks, 24(9):933 -- 942, 2011. [6] Chunlin Ji, Daniel Merl, Thomas B Kepler, and Mike West. Spatial mixture modelling for unobserved point processes: Examples in immunofluorescence histology. Bayesian analysis (Online), 4(2):297, 2009. [7] Athanasios Kottas and Bruno Sans´o. Bayesian mixture modeling for spatial poisson process intensities, with applications to extreme value analysis. Journal of Statistical Planning and Inference, 137(10):3151 -- 3163, 2007. [8] Alex Kulesza and Ben Taskar. Determinantal point processes for machine learning. Machine Learning, 5(2- 3):123 -- 286, 2012. [9] Ed S Lein, Michael J Hawrylycz, Nancy Ao, Mikael Ayres, Amy Bensinger, Amy Bernard, Andrew F Boe, Mark S Boguski, Kevin S Brockway, Emi J Byrnes, et al. Genome-wide atlas of gene expression in the adult mouse brain. Nature, 445(7124):168 -- 176, 2007. [10] Henry Markram, Maria Toledo-Rodriguez, Yun Wang, Anirudh Gupta, Gilad Silberberg, and Caizhi Wu. In- terneurons of the neocortical inhibitory system. Nat Rev Neurosci, 5(10):793 -- 807, October 2004. 9 [11] Alexander Smola and Shravan Narayanamurthy. An architecture for parallel topic models. Proceedings of the VLDB Endowment, 3(1-2):703 -- 710, 2010. [12] Amit Zeisel, Ana B Munoz-Manchado, Simone Codeluppi, Peter Lonnerberg, Gioele La Manno, Anna Jur´eus, Sueli Marques, Hermany Munguba, Liqun He, Christer Betsholtz, et al. Cell types in the mouse cortex and hippocampus revealed by single-cell rna-seq. Science, 347(6226):1138 -- 1142, 2015. 10 Appendix for Discovering Neuronal Cell Types and Their Gene Expression Profiles Using a Spatial Point Process Mixture Model A Morphological Basis Extraction We aim to characterize the morphological basis for all cells with different size, orientation, expression profiles and spatial distribution. The traditional sparse coding introduces too many free parameters and is not suitable for compact morphological basis learning. We instead propose Gaussian prior convolutional sparse coding (GPCSC). The intuition for using convolution is due to the frequent replication of cells of similar shapes and the translation invariance property. Traditional sparse coding would learn both the shape of the cell and the location of the cell. But the convolutional sparse coding would only learn the shape here. We characterize cell spatial distribution via decoding the sparse activation map. To formulate the problem formally: let I be the image observed, then the convolutional sparse coding model generates observed image I using filters (resembling cell shapes)F superposed at locations indicated by the activation map M (whose sparsity pattern indicates cell spatial distribution and activation amplitude indicates gene expression profiles. ) Our goals of segmenting cells, extracting cell basis, and estimating gene profiles and cell locations are reduced to this optimization learning problem: min Fm,M n m(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xn 2 F m(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) I n − kXm=1 Fm ⋆ M n +Xn Xm λ kM n mk0 , s.t. Fm(x, y) ≥ 0, kFmk2 F = 1, M (n) m (x, y) ≥ 0. (5) where I n is the nth image associated with the gene we are interested in with Dx × Dy pixels, i.e., I n ∈ RD×D. We call the Fm ∈ Rd×d filter, where d is set to capture the local cell morphological information. The spatial m ∈ R(D−d+1)×(D−d+1) which represents the position of the filter Fm being m(x, y) = 1, then Fm is active at I n(x : x + d − 1, y : y + d − 1). coefficient for image I n is denoted as H (n) active on image I n. More precisely, if H n A.1 Gaussian Prior Convolutional Sparse Coding The popular alternating approach between matching pursuit to learn activation map M and k-SVD to learn F is general applicable to any object detection problem in image processing. However, this approach causes inexact cell number estimation as filters with multi-modality (i.e., multiple cells) are learnt. We resolve this issue by proposing an Gaussian probability density function prior on the filters to guarantee single cell detection and achieve accurate cell number estimation. The support of M is also limited to the local maxima indicating cell centers. Note that our cell are not donut shaped, and it is reasonable to assume the darkest point being the cell center. Therefore, we optimize over the objective min kPn I n −Pm Fm ⋆ M n solving the optimization problem. If we define the residual asPn I n−PnPm bFm ⋆cM n mk0 such that Fm are 2 − D Gaussian densities with priori set top 2 principal radius and orientation. Alternating Minimization is used to m, the gradient of the objective reduced to an iterative approach of updating filters, compute residual, optimizing activation map based on residual, (i, j) are convolution of the compute residual and updating filters again. It is easy to see that both ∂L ∂Fm residual and the other variable rotated by angle π. 2 +PnPm λ kM n (i, j) and ∂L ∂Hm mk2 A.2 Image Registration/Alignment A structure represents a neuronanatomical region of interest. Structures are grouped into ontologies and organized in a hierarchy or structure graph. We are interested in the somatosensory cortex area. So we use the affine transform from Allen Brain Institute [1, 9] to align all the in-situ hybridization images with the Atlas brain to extract the correct region. 11
1805.11704
2
1805
2018-06-29T20:17:16
Deep Semantic Architecture with discriminative feature visualization for neuroimage analysis
[ "q-bio.NC", "cs.AI", "q-bio.QM" ]
Neuroimaging data analysis often involves \emph{a-priori} selection of data features to study the underlying neural activity. Since this could lead to sub-optimal feature selection and thereby prevent the detection of subtle patterns in neural activity, data-driven methods have recently gained popularity for optimizing neuroimaging data analysis pipelines and thereby, improving our understanding of neural mechanisms. In this context, we developed a deep convolutional architecture that can identify discriminating patterns in neuroimaging data and applied it to electroencephalography (EEG) recordings collected from 25 subjects performing a hand motor task before and after a rest period or a bout of exercise. The deep network was trained to classify subjects into exercise and control groups based on differences in their EEG signals. Subsequently, we developed a novel method termed the cue-combination for Class Activation Map (ccCAM), which enabled us to identify discriminating spatio-temporal features within definite frequency bands (23--33 Hz) and assess the effects of exercise on the brain. Additionally, the proposed architecture allowed the visualization of the differences in the propagation of underlying neural activity across the cortex between the two groups, for the first time in our knowledge. Our results demonstrate the feasibility of using deep network architectures for neuroimaging analysis in different contexts such as, for the identification of robust brain biomarkers to better characterize and potentially treat neurological disorders.
q-bio.NC
q-bio
Deep Semantic Architecture with discriminative feature visualization for neuroimage analysis Arna Ghosh Integrated Program in Neuroscience McGill University Montréal, QC H3A 0G4. Canada [email protected] Fabien Dal Maso École de kinésiologie et des sciences de l'activité physique Université de Montréal Montréal, QC H3T 1J4. Canada [email protected] Marc Roig School of Physical and Occupational Therapy McGill University Montréal, QC H3A 0G4. Canada. [email protected] Georgios D Mitsis Department of Bioengineering McGill University Montréal, QC H3A 0G4. Canada. [email protected] Marie-Hélène Boudrias School of Physical and Occupational Therapy McGill University Montréal, QC H3A 0G4. Canada. [email protected] Abstract Neuroimaging data analysis often involves a-priori selection of data features to study the underlying neural activity. Since this could lead to sub-optimal feature selection and thereby prevent the detection of subtle patterns in neural activity, data- driven methods have recently gained popularity for optimizing neuroimaging data analysis pipelines and thereby, improving our understanding of neural mechanisms. In this context, we developed a deep convolutional architecture that can identify discriminating patterns in neuroimaging data and applied it to electroencephalogra- phy (EEG) recordings collected from 25 subjects performing a hand motor task before and after a rest period or a bout of exercise. The deep network was trained to classify subjects into exercise and control groups based on differences in their EEG signals. Subsequently, we developed a novel method termed the cue-combination for Class Activation Map (ccCAM), which enabled us to identify discriminating spatio-temporal features within definite frequency bands (23–33 Hz) and assess the effects of exercise on the brain. Additionally, the proposed architecture allowed the visualization of the differences in the propagation of underlying neural activity across the cortex between the two groups, for the first time in our knowledge. Our results demonstrate the feasibility of using deep network architectures for neuroimaging analysis in different contexts such as, for the identification of robust brain biomarkers to better characterize and potentially treat neurological disorders. Preprint. Work in progress. 1 Introduction Skilled motor practice facilitates the formation of an internal model of movement, which may be later used to anticipate task specific requirements. These internal models are more susceptible to alterations during and immediately following practice and become less susceptible to alterations over time, a process called consolidation [5, 23]. A single bout of cardiovascular exercise, performed in close temporal proximity to a session of skill practice, has shown to facilitate motor memory consolidation [24]. Several potential mechanisms underlying the time-dependent effects induced by acute exercise on motor memory consolidation have been identified, such as increased availability of neurochemicals [28] and increased cortico-spinal excitability [19]. However, the distinct contribution of specific brain areas and the precise neurophysiological mechanisms underlying the positive effects of acute cardiovascular exercise on motor memory consolidation remain largely unknown. Electroencephalography (EEG) is a popular technique used to study the electrical activity from different brain areas. The EEG signal arises from synchronized postsynaptic potentials of neurons that generate electrophysiological oscillations in different frequency bands. During movement, the EEG signal power spectrum within the alpha (8–12 Hz) and beta (15–29 Hz) range decreases in amplitude and this is thought to reflect increased excitability of neurons in sensorimotor areas [7, 18, 20, 25]. This phenomenon is termed Event-Related Desynchronization (ERD). Alpha- and beta-band ERD have been shown to be modulated during motor skill learning in various EEG studies [4, 14, 34]. There is converging evidence suggesting an association of cortical oscillations in the motor cortex with neuroplasticity events underlying motor memory consolidation [4, 22]. In this context, our aim was to study the add-on effects of exercise on motor learning in terms of modulation of EEG-based ERD. Many neuroimaging studies, including EEG ones, rely on the a-priori selection of features from the recorded time-series. This could lead to sub-optimal feature selection and could eventually prevent the detection of subtle discriminative patterns in the data. Alternatively, data-driven approaches such as deep learning allow discovery of the optimal discriminative features in a given dataset. Convolutional Neural Networks (CNNs) and Recurrent Neural Networks (RNNs) have been applied to computer vision and speech processing datasets [17, 12, 32, 15] with great success. They have also been used successfully in the neuroimaging domain to learn feature representations for Magnetic Resonance Image segmentations [21] and EEG data decoding [2, 26, 30] among others. Most studies using CNNs for EEG have been restricted to the classification of EEG data segments into known categories. However, the usefulness of CNNs to improve our understanding of the underlying neural bases is less straightforward, primarily due to the difficulty into visualizing and interpreting the feature space learnt by the CNN. Our work addresses existing caveats in applying deep learning architectures, such as CNNs, to analyzing EEG data by means of three novel contributions – 1. We used two parallel feature extraction streams to discover informative features from EEG data before and after an intervention and subsequently characterize the modulatory effect on these derived features rather than on the raw EEG data itself; 2. We incorporated a subject prediction adversary component in the network architecture to learn subject-invariant, group-related features instead of subject-specific features; 3. We developed a novel method, termed cue-combination for Class Activation Map (ccCAM), to visualize the features extracted by the CNN after training We used this CNN-based deep network architecture to identify exercise-induced changes in neural activity from EEG signals recorded during an isometric motor task. The training was carried out in a hierarchical structure – first for time-frequency and then for topographical data maps. Visualizing the features after each stage of the training allowed us to identify frequency bands and the corresponding brain regions that were modulated by the add-on effects of acute exercise on motor learning. The majority of previous related studies have leveraged large-scale datasets consisting of hundreds of subjects for training purposes, which may be a limiting factor for applying powerful deep learning methodologies to data of smaller sample size. Therefore, one of our aims was to develop a method that can be used both for small-scale and large-scale studies. To this end, we added a regularizer that prevented the feature extraction part of the CNN from learning subject-specific features, thus promoting the identification of group-specific features only. 2 2 Dataset The dataset used in this work consisted of EEG recordings from 25 healthy subjects. The experiment and data collection are detailed elsewhere [8]. Briefly, 25 right-handed healthy subjects were recruited and assigned to the Control (CON, n=13 subjects) or Exercise (EXE, n=12 subjects) groups. Figure 1: Illustration of Experiment Protocol. Each subject reported to the laboratory on four occasions as shown in Figure 1. Visit 1 required the participants to go through a Graded Exercise Test (GXT), which was used to determine their cardiorespiratory fitness. Visit 2 was conducted at least 48 hrs after the GXT to avoid potential long- term effects of exercise on memory [3, 13]. EEG recordings were collected at baseline while subjects performed isometric handgrips, which corresponded to 50 repetitions of visually cued isometric handgrips with their dominant right hand using a hand clench dynamometer (Biopac, Goleta, CA, USA). Each contraction was maintained for 3.5 sec at 15% of each participant's maximum voluntary contraction (MVC) and followed by a 3 to 5 sec rest period. The baseline assessment was followed by the practice of a visuo-motor tracking task (skill acquisition), which was used for the calculation of the motor learning score. Participants were then randomly assigned to two groups. The exercise group (EXE) performed a bout of high-intensity interval cycling of 15 min, while the control group (CON) rested on the cycle ergometer for the same amount of time. The same EEG recordings collected at baseline were repeated 30, 60 and 90 min after the exercise or rest period.EEG activity was recorded using a 64-channel ActiCap cap (BrainVision, Munich, Germany) with electrode locations arranged according to the 10–20 international system. Electrical conductive gel was inserted at each electrode site to keep impedances below 5 kΩ. EEG signals were referenced to the FCz electrode and sampled at 2500 Hz. 3 Methods The analysis pipeline was first applied to the time and frequency domain data without incorporating spatial information. Subsequently, it was applied to the data obtained by creating topographical maps corresponding to distribution of activity in specific frequency bands across the cortex. The entire pipeline consisted of 3 segments , i.e.– Preprocessing, CNN training and ccCAM generation. 3.1 Time-Frequency (TF) maps 3.1.1 Preprocessing EEG data preprocessing was similar to that performed in a previous study [8] and was performed using the Brainstorm Matlab toolbox [29]. EEG signals were bandpass-filtered between 0.5 Hz and 55 Hz and average-referenced. Continuous data were visually inspected and portions of signals with muscle or electrical transient artifacts were rejected. Independent component analysis (ICA) was subsequently applied on each dataset (total number of components: 20) and between one and two eye-blink related components were rejected based on their topography and time signatures [9]. The resulting dataset was epoched with respect to the period of time (3.5 sec) corresponding to the appearance of the visual cue that triggered the initiation of the isometric handgrips (n = 50/subject). Finally, each trial was visually inspected and those containing artifacts were manually removed. Morlet wavelet (wave number = 7) coefficients between 1 to 55 Hz with 1 Hz resolution were extracted to obtain time-frequency decompositions of the EEG data. The time-frequency data for 3 each electrode were consequently normalized with respect to their spectral power before the start of the grip event, as calculated from a window of 0.5 sec. Following this, an average over all trials was calculated in order to obtain a single time-frequency map for each electrode. Further steps were applied on the EEG recording segment corresponding to 0.5–3.5 sec after the appearance of the visual cue, i.e. during the handgrip task, to perform the subsequent analysis. 3.1.2 CNN training t and As t and As The overall CNN architecture that we developed is shown in Figure 2. Following preprocessing of the data, time-frequency maps for each electrode and session – at baseline and 90 min after exercise or a rest period (post-intervention session) – were obtained. The data for each session was then rearranged to form 2D matrices comprising of the frequency spectra for all electrodes at a given time instant t. Each matrix had a dimension of 64 × 55 (64 electrodes × 55 frequency bands). For training the network, a pair of matrices was used – the first corresponding to time point t from the baseline session and the second corresponding to the same time point t from the post-intervention session. Each pair was labeled as either exercise or control, depending on the group allocation. Structuring the data in this fashion allowed the network to take into account the inter-subject variability in baseline measures and therefore did not require the experimenter to adopt techniques for normalizing the EEG signal from the post-intervention session with respect to the baseline session. Thus, the network was expected to capture the EEG features that were modulated by the add-on effects of acute exercise. Dataset Notation:- B and A represent the entire data tensor at baseline and post-intervention respec- tively. Each data tensor consists of data matrices from all 25 subjects and timepoints. For subject s, the goal was to classify whether the tuple containing the matrices Bs t (where t denotes timepoint) belongs to the EXE or CON groups. To this end, we used a deep convolutional network that was optimized for the task. The network architecture is similar to the one used in [1]. Features from matrices Bs t were extracted using a network termed the Base CNN. The difference between the obtained feature vectors was passed to a discriminator network, termed the Top NN, to predict the correct group in which each pair belongs to. The schematic view of the architecture is shown in Figure 2 and details of each network's architecture are provided in Tables S1 and S2 in supplementary material respectively. Since the sampling frequency was 2500 Hz and the time period of interest was 3 sec long, t ∈ [1, 7500] for each subject s. The convolutions performed in the Base CNN were with respect to the frequency domain and not the electrode (sensor) domain. This is because the former was laid out in a semantic order of increasing frequencies, as opposed to the latter, which was not arranged by the spatial locations of the electrodes. Consequently, we expected the features extracted by the Base CNN to be the frequency bands significantly affected by exercise. Therefore, all convolutional filters in the Base CNN were implemented as 1 × n 2D filters, where n is the extent of the filter in the frequency domain. The same holds for the Max-Pooling layers. Initially, a network that did not include an adversary loss component (Figure S1.a from Supplementary material) was used; however, it was found that this network was able to learn subject-specific features as opposed to subject-invariant, exercise-related features. This is illustrated in Figure S2 (Supplementary material) and Table 1. In most neuroimaging studies, the number of participants scanned is limited, which typically restricts deep networks from learning subject-invariant features. To address this issue, we followed a domain adaptation approach. Specifically, each subject was considered as a separate domain comprising of subject-specific features along with subject-invariant, exercise-related features. Since our goal was to learn features mainly related to the effect of exercise on the consolidation of motor memory, we incorporated the domain confusion strategy [31] to train the network, thus adding the subject discriminator as an adversary (Figure 2 – bottom right). Specifically, we added this network in parallel to the Top NN with similar model capacity (see Table S3 in supplementary material for architecture details). Network Architecture Notation:- The feature extractor operation and parameters of the Base CNN are denoted as fθf and θf respectively, the Top NN feature discrimination operator and its parameters are denoted by hθt and θt respectively, while the subject discrimination operator and its parameters are denoted by hθs and θs respectively. The input tuple is denoted by x and its corresponding group and subject labels by yg and ys respectively. We used the Negative Log Likelihood (NLL) loss for each classifier with the Adam optimizer [16] in Torch [6] for training the network. The Subject 4 Figure 2: Modified deep network architecture with an adversary component (bottom right) to avoid subject discrimination. The adversary is a novel addition to enable the use of CNNs to smaller-scale neuroimaging studies with a limited number of subjects. Dimensions corresponding to TF maps are shown here. Discriminator was trained to minimize the subject prediction loss given by – Js(θs, θf ) = −[ m X i=1 logh(y(i) s ) θs (fθf (x(i)))] The Top NN was trained to minimize the group prediction loss given by – Jg(θt, θf ) = −[ m X i=1 logh (y(i) g ) θt (fθf (x(i)))] (1) (2) We trained the feature extractor, Base CNN, in a manner such that the features extracted would be agnostic to the originating subject, therefore, the target distribution for the subject prediction network was a uniform distribution. Hence, we used the domain confusion loss [31] over the gradient reversal layer [11] and used the Kullback-Leibler (KL) divergence from the uniform distribution over 25 classes (25 subjects) as our loss metric. Conclusively, the Base CNN was trained to minimize the loss given by – Jf (θf , θt, θs) = −[ m X i=1 logh (y(i) g ) θt (fθf (x(i)))] + λ[ m X i=1 KL(U, hθs (fθf (x(i)))] (3) where KL(P, Q) denotes the KL divergence between distributions P & Q, U denotes the uniform distribution, m denotes the total number of training examples and λ is a hyperparamater that deter- mines the weight for the subject discrimination regularizer. Here, we used a 80-20 split of the data set, whereby 80% was used for training and 20% was used for validation. ccCAM 3.1.3 A major contribution of the present work is the development of a novel method for the visualization of the features that guide the proposed network's decision. Although well-known techniques used in the computer vision literature include the use of Global Average Pooling (GAP) [33] and grad-CAM [27], these methods are not suited for the neuroscience paradigm considered here. For instance, GAP requires averaging the activations of each filter map, i.e. each channel of the extracted feature tensor. This leads to loss of information related to electrode positions, as convolutions were performed only in the frequency domain. Specifically, we applied GAP and grad-CAM to our data and we were unable to obtain adequate classification accuracy (≈ 56%) with a GAP layer in the network. Also, grad-CAM is sensitive to absolute scale of the features in the input data and hence yielded results that were biased towards frequency bands with higher power-values, namely the lower frequency bands (<10 Hz). 5 Given these limitations in existing analytic methods, we used the linear cue-combination theory used in human perception studies [10] to develop a method that explains the network's decisions. Let us consider for example, a CNN with only 2 channels , i.e. filter maps, in the final feature tensor extracted after convolutions. Each of these filter maps preserve the spatial and/or semantic structure of the input data. Each of these filter maps acts as a "cue" to the network's classifier layers, denoted as c1 and c2. If we denote the desired class label as y1 and assuming c1 and c2 to be independent to each other, we can use Bayes' Theorem to write – P (y1c1, c2) = P (c1, c2y1)P (y1) P (c1, c2) = P (c1y1)P (c2y1)P (y1) P (c1)P (c2) = P (y1c1)P (y1c2) P (y1) (4) If the likelihood for predicting y1 due to cue ci is Gaussian with mean µi and variance σ2 i , the maximum likelihood estimate (MLE) yields the combined cue, denoted by c∗, that summarizes the important features on which the network bases its decisions. Therefore, the combined cue, c∗, is the desired Class Activation Map (CAM). c∗ = 2 X i=1 wiµi where wi = 1/σ2 i 2 1/σ2 i P i=1 Since the network is trained, µi = ci. To calculate the values of σi, we used the NLL loss values. The NLL loss with a cue removed provided an estimate of the σ associated with that cue. (5) (6) (7) ǫ = −logP (y1c1, c2) = −logP (y1c1) − logP (y1c2) + logP (y1) [From eq 4] ǫ1 = ǫc1=0 = −logP (y1c1 = 0) − logP (y1c2) + logP (y1) ǫ1 − ǫ = logP (y1c1) − logP (y1c1 = 0) µ2 1 2σ2 1 2(ǫ1 − ǫ) µ2 1 = = 21 1 σ Therefore, Using the estimated σi, the CAM corresponding to the correct class for each input was generated. Since in the present case µi corresponds to a 2D matrix, the denominator in Equation 7 was replaced by the mean-squared value of the corresponding matrix. The obtained CAMs were subsequently group-averaged to extract frequency bands that contain features characteristic to each group (CON and EXE). 3.2 Topographical maps Topographical maps were created using the frequency bands obtained from the ccCAM corresponding to the TF-maps. The average power within each frequency band for all electrodes at time point t was used to construct a 64 × 64 matrix by projecting the average power value of each electrode to a point corresponding to the its spatial position. Since this procedure yielded a sparse matrix, cubic interpolation was used to obtain a continuous image depicting the distribution of activity within each frequency band over the entire head. A total of three such matrices were packed together to form a 3 × 64 × 64 tensor corresponding to activity maps at times t, t + 1 and t + 2 respectively. The entire data tensor for a given subject was created by taking non-overlapping time windows. Hence, the total number of tensors for each subject was equal to 2500. Similar to the analysis of TF-maps, we trained a CNN-based network to classify each data tensor into the CON and EXE groups. Since the inputs were 2D image tensors here, we used 2D convolutional 6 [µ2 j ](i) m X i=1 2j 1 σ σi is estimated over the entire dataset as shown in Equation 7. 2[(ǫj − ǫ)](i) = filters in the Base CNN (see Tables S4, S5 and S6 from supplementary material for more details). Following training, ccCAM was applied to obtain CAMs for each subject at each time instant during the task execution. 4 Results and Discussion The results presented here illustrate the differences between the Baseline and 90 min post exercise/rest datasets. The network architecture details for each type of data (TF and Topographical) map are presented in the supplementary material, along with details regarding the chosen hyperparameters. 4.1 Time-Frequency maps We observed that the features extracted by Base CNN, without any subject prediction regularizer, could be used to identify the subject corresponding to any given data tensor. As the subject discriminator regularization was given more weight by increasing λ, the Base CNN learned to extract features that were agnostic to the originating subject. However for very high λ values, the extracted features could not be used to discriminate the EXE and CON groups, suggesting that the Base CNN was unable to learn discriminative features. The loss values obtained post-training for four different values of λ are shown in Table 1. The choice of an optimal value for λ depends on two factors – group prediction accuracy and subject prediction accuracy. To identify subject-invariant features, we aimed to obtain an optimal value of λ that achieved good group prediction accuracy but poor subject prediction accuracy. Consequently, this required a good tradeoff between the two prediction accuracies. Table 1: Variation of Loss values with λ after training network on TF maps. Group prediction loss Subject prediction loss (NLL) ≈ 0 ≈ 1.5 ≈ 2.6 ≈ 3.2 KL divergence loss from Uniform distribtion ≈ 0.3 ≈ 0.07 ≈ 0.004 ≈ 0.0002 λ 0 10 13 15 (NLL) ≈ 0 ≈ 0.1 ≈ 0.4 ≈ 0.68 According to this procedure, the model corresponding to λ = 13 was used for ccCAM generation. The average loss over a batch for subject prediction was around 2.6, which roughly predicted the 13. The group prediction accuracy was 99.984% (99.969% for correct subject with a confidence of 1 CON and 100% for EXE). Hence the extracted features achieved excellent group prediction, while all subjects in the group were predicted with roughly equal probability (CON and EXE consisted of 13 and 12 subjects respectively). The ccCAM obtained is shown in Figure 3. Figure 3: Time-averaged TF map ccCAM averaged over electrodes & subjects showing "discrimina- tive" frequencies. As one of the main goals of this study was to identify the frequency bands that contained significant information, we calculated the ccCAMs for all timepoints and then group-averaged (averaging across all timepoints and subjects in a group) the maps to get two 2D maps – for the CON and EXE groups. We plotted the average activation within each frequency band in each of these 2D maps to obtain the plots in Figure 3. The bold lines denote the group-mean and the shaded regions span 1 standard 7 error over all subjects in the group. The two plots are significantly different within the band 23–33 Hz. This band lies within the wider beta-band and agrees with findings in [8] where beta-band desynchronization was found to be significantly modulated by exercise. It is important to note that the ccCAM highlights the differences between the 90min and baseline EEG recordings. The negative values in a frequency band indicate that the ERD was smaller after than before the exercise. This also agrees with findings in [8] and implies that decreased neural activity was required to perform the hand-grip task after exercise. The p-value calculated from the ccCAM outputs within this frequency band was equal to 0.021, while the corresponding p-value from the original time-frequency data tensor was equal to 0.0134. This suggests that had the band of interest in the previous study [8] been chosen to be 23–33 Hz, instead of the wider beta-band (15–29 Hz), similar, statistically significant inferences would have been drawn. Topographical maps were created to understand the distribution of the activity within the 23–33 Hz frequency band over the cortex. After training a network to classify into the CON and EXE groups from topographical maps, a classification accuracy of 98.70% (98.94% for CON and 98.43% for EXE) was obtained for λ = 5. Generating ccCAMs for the topographical maps revealed the propagation of the discriminative activity across the cortex. A video showing this traveling property of this activity is included in the supplementary material. Some snapshots from the video are shown in Figure 4. Figure 4: Topographical map ccCAM averaged over subjects in a group showing regions with difference in activity before and 90min after rest/exercise. Top row shows CON group. Bottom row shows EXE group for a brief time window, where a streak of activity can be seen to move across time from the parietal to motor cortex. To the best of our knowledge, this traveling pattern of activity across the cortex while performing an isometric handgrip has not been demonstrated before. These oscillations could allow us to visualize the neural mechanisms involved in maintaining a constant grip-force output. Further investigation into the correlation of these activities with the observed error signal while performing the task is required to understand these mechanisms more precisely. As expected, differences in the ccCAM of EXE group before and after exercise were higher in magnitude as compared to those in the CON group (see Figure S4 in supplementary material), thus indicating the modulatory effects of an acute bout of high-intensity exercise. 5 Conclusion This work introduces a deep learning architecture for the analysis of EEG data and shows promising results in terms of discriminating the participants that underwent an acute bout of high-intensity exercise/rest in close temporal proximity to performing a motor learning task. Importantly, the proposed novel method enabled us to visualize the features learnt by deep networks such as CNNs, which may in turn yield better interpretation of their classification basis. The results are in general agreement with those reported in a previous study using more standard statistical analysis for a-priori selected features on the same dataset [8], with our analysis revealing a narrower, more-specific frequency band associated with exercise-induced changes. In addition, our method revealed, for the first time, the traveling pattern of cortical activity while subjects were performing isometric handgrips. 8 Therefore, our approach demonstrates scope of identifying discriminative features in a completely data-driven manner. The proposed method is not restricted to the EEG modality and dataset described here. Hence, it paves the way for applying equivalent deep learning methods to datasets obtained from neuroimaging studies of differing scales and varying modalities (eg. magnetoencephalography – MEG). This, in turn, yields great potential to accelerate research oriented towards identification of neurophysiological changes associated with various neurological disorders and ultimately lead to design of optimized and individualized intervention strategies. References [1] P. Agrawal, J. Carreira, and J. Malik. Learning to see by moving. In Computer Vision (ICCV), 2015 IEEE International Conference on, pages 37–45. IEEE, 2015. [2] P. Bashivan, I. Rish, M. Yeasin, and N. Codella. Learning representations from eeg with deep recurrent-convolutional neural networks. arXiv preprint arXiv:1511.06448, 2015. [3] N. Berchtold, G. Chinn, M. Chou, J. Kesslak, and C. Cotman. Exercise primes a molecu- lar memory for brain-derived neurotrophic factor protein induction in the rat hippocampus. Neuroscience, 133(3):853–861, 2005. [4] T. W. Boonstra, A. Daffertshofer, M. Breakspear, and P. J. Beek. Multivariate time–frequency analysis of electromagnetic brain activity during bimanual motor learning. Neuroimage, 36(2):370–377, 2007. [5] T. Brashers-Krug, R. Shadmehr, and E. Bizzi. Consolidation in human motor memory. Nature, 382(6588):252, 1996. [6] R. Collobert, K. Kavukcuoglu, and C. Farabet. Torch7: A matlab-like environment for machine learning. In BigLearn, NIPS workshop, number EPFL-CONF-192376, 2011. [7] N. E. Crone, D. L. Miglioretti, B. Gordon, J. M. Sieracki, M. T. Wilson, S. Uematsu, and R. P. Lesser. Functional mapping of human sensorimotor cortex with electrocorticographic spectral analysis. i. alpha and beta event-related desynchronization. Brain: a journal of neurology, 121(12):2271–2299, 1998. [8] F. Dal Maso, B. Desormeau, M.-H. Boudrias, and M. Roig. Acute cardiovascular exercise promotes functional changes in cortico-motor networks during the early stages of motor memory consolidation. NeuroImage, 174:380–392, 2018. [9] A. Delorme and S. Makeig. Eeglab: an open source toolbox for analysis of single-trial eeg dynamics including independent component analysis. Journal of neuroscience methods, 134(1):9–21, 2004. [10] M. O. Ernst and M. S. Banks. Humans integrate visual and haptic information in a statistically optimal fashion. Nature, 415(6870):429, 2002. [11] Y. Ganin, E. Ustinova, H. Ajakan, P. Germain, H. Larochelle, F. Laviolette, M. Marchand, and V. Lempitsky. Domain-adversarial training of neural networks. The Journal of Machine Learning Research, 17(1):2096–2030, 2016. [12] A. Graves, A.-r. Mohamed, and G. Hinton. Speech recognition with deep recurrent neural net- works. In Acoustics, speech and signal processing (icassp), 2013 ieee international conference on, pages 6645–6649. IEEE, 2013. [13] M. E. Hopkins, F. C. Davis, M. R. VanTieghem, P. J. Whalen, and D. J. Bucci. Differential effects of acute and regular physical exercise on cognition and affect. Neuroscience, 215:59–68, 2012. [14] S. Houweling, A. Daffertshofer, B. W. van Dijk, and P. J. Beek. Neural changes induced by learning a challenging perceptual-motor task. Neuroimage, 41(4):1395–1407, 2008. [15] A. Karpathy, G. Toderici, S. Shetty, T. Leung, R. Sukthankar, and L. Fei-Fei. Large-scale video classification with convolutional neural networks. In Proceedings of the IEEE conference on Computer Vision and Pattern Recognition, pages 1725–1732, 2014. 9 [16] D. P. Kingma and J. Ba. Adam: A method for stochastic optimization. arXiv preprint arXiv:1412.6980, 2014. [17] A. Krizhevsky, I. Sutskever, and G. E. Hinton. Imagenet classification with deep convolutional neural networks. In Advances in neural information processing systems, pages 1097–1105, 2012. [18] C. Neuper and G. Pfurtscheller. Event-related dynamics of cortical rhythms: frequency-specific features and functional correlates. International journal of psychophysiology, 43(1):41–58, 2001. [19] F. Ostadan, C. Centeno, J.-F. Daloze, M. Frenn, J. Lundbye-Jensen, and M. Roig. Changes in corticospinal excitability during consolidation predict acute exercise-induced off-line gains in procedural memory. Neurobiology of learning and memory, 136:196–203, 2016. [20] G. Pfurtscheller, B. Graimann, J. E. Huggins, S. P. Levine, and L. A. Schuh. Spatiotemporal patterns of beta desynchronization and gamma synchronization in corticographic data during self-paced movement. Clinical neurophysiology, 114(7):1226–1236, 2003. [21] S. M. Plis, D. R. Hjelm, R. Salakhutdinov, E. A. Allen, H. J. Bockholt, J. D. Long, H. J. Johnson, J. S. Paulsen, J. A. Turner, and V. D. Calhoun. Deep learning for neuroimaging: a validation study. Frontiers in neuroscience, 8:229, 2014. [22] B. Pollok, D. Latz, V. Krause, M. Butz, and A. Schnitzler. Changes of motor-cortical oscillations associated with motor learning. Neuroscience, 275:47–53, 2014. [23] E. M. Robertson, A. Pascual-Leone, and R. C. Miall. Current concepts in procedural consolida- tion. Nature Reviews Neuroscience, 5(7):576, 2004. [24] M. Roig, S. Nordbrandt, S. S. Geertsen, and J. B. Nielsen. The effects of cardiovascular exercise on human memory: a review with meta-analysis. Neuroscience & Biobehavioral Reviews, 37(8):1645–1666, 2013. [25] R. Salmelin, M. Hámáaláinen, M. Kajola, and R. Hari. Functional segregation of movement- related rhythmic activity in the human brain. Neuroimage, 2(4):237–243, 1995. [26] R. T. Schirrmeister, J. T. Springenberg, L. D. J. Fiederer, M. Glasstetter, K. Eggensperger, M. Tangermann, F. Hutter, W. Burgard, and T. Ball. Deep learning with convolutional neural networks for eeg decoding and visualization. Human brain mapping, 38(11):5391–5420, 2017. [27] R. R. Selvaraju, M. Cogswell, A. Das, R. Vedantam, D. Parikh, and D. Batra. Grad-cam: Visual explanations from deep networks via gradient-based localization. See https://arxiv. org/abs/1610.02391 v3, 7(8), 2016. [28] K. Skriver, M. Roig, J. Lundbye-Jensen, J. Pingel, J. W. Helge, B. Kiens, and J. B. Nielsen. Acute exercise improves motor memory: exploring potential biomarkers. Neurobiology of learning and memory, 116:46–58, 2014. [29] F. Tadel, S. Baillet, J. C. Mosher, D. Pantazis, and R. M. Leahy. Brainstorm: a user-friendly application for meg/eeg analysis. Computational intelligence and neuroscience, 2011:8, 2011. [30] P. Thodoroff, J. Pineau, and A. Lim. Learning robust features using deep learning for automatic seizure detection. In Machine Learning for Healthcare Conference, pages 178–190, 2016. [31] E. Tzeng, J. Hoffman, T. Darrell, and K. Saenko. Simultaneous deep transfer across domains and tasks. In Computer Vision (ICCV), 2015 IEEE International Conference on, pages 4068–4076. IEEE, 2015. [32] X. Zhang and Y. LeCun. Text understanding from scratch. arXiv preprint arXiv:1502.01710, 2015. [33] B. Zhou, A. Khosla, A. Lapedriza, A. Oliva, and A. Torralba. Learning deep features for discriminative localization. In Computer Vision and Pattern Recognition (CVPR), 2016 IEEE Conference on, pages 2921–2929. IEEE, 2016. 10 [34] P. Zhuang, C. Toro, J. Grafman, P. Manganotti, L. Leocani, and M. Hallett. Event-related desynchronization (erd) in the alpha frequency during development of implicit and explicit learning. Electroencephalography and clinical neurophysiology, 102(4):374–381, 1997. 11 Deep Semantic Architecture with discriminative feature visualization for neuroimage analysis Arna Ghosh Integrated Program in Neuroscience McGill University Montréal, QC H3A 0G4. Canada [email protected] Fabien Dal Maso École de kinésiologie et des sciences de l'activité physique Université de Montréal Montréal, QC H3T 1J4. Canada [email protected] Marc Roig School of Physical and Occupational Therapy McGill University Montréal, QC H3A 0G4. Canada. [email protected] Georgios D Mitsis Department of Bioengineering McGill University Montréal, QC H3A 0G4. Canada. [email protected] Marie-Hélène Boudrias School of Physical and Occupational Therapy McGill University Montréal, QC H3A 0G4. Canada. [email protected] Supplementary Material 1 Network Architecture Notation:- Conv denotes the 2D Spatial Convolutional layer. ReLU denotes the Rectified Linear Unit Layer that adds non-linearity to the network. MaxPool denotes the 2D Spatial Max Pooling layer. FullyConn denotes a Fully Connected layer, also known as the linear layer of the network. 1.1 TF maps Layer 0 1 2 3 4 5 6 Type Input Conv ReLU Conv ReLU MaxPool MaxPool Maps and Neurons 1M × 64 × 55N 6M × 64 × 28N 6M × 64 × 28N 6M × 64 × 14N 16M × 64 × 14N 16M × 64 × 14N 16M × 64 × 7N Filter Size - 1×5 - 1×2 1×5 - 1×2 Table S1: Network architecture used for EEG feature extraction network (Base CNN). The output of the network is a tensor of dimensions 16 × 64 × 7. Preprint. Work in progress. (a) Basic Architecture without adversary (b) Modified Architecture with adversary to avoid subject discrimination Figure S1: Deep Network Architecture. The initial choice of architecture (without any adversary) gives good subject prediction accuracy from features extracted by the Base CNN. Therefore, a subject discriminator of roughly the same model capacity as the Top NN is added. The subject discrimination acts as a regularizer while training and avoids the Base CNN from learning subject specific features. Type Input Flatten Maps and Neurons 16M × 64 × 7N 7168N Filter Size Layer 0 1 2 3 4 5 Layer 0 1 2 3 4 5 Dropout (p=0.5) FullyConn ReLU FullyConn Dropout (p=0.5) FullyConn ReLU FullyConn - 8N 8N 2N - 8N 8N 25N 2 - - - 1×1 - 1×1 - - - 1×1 - 1×1 Table S2: Network architecture used for group discrimination network (Top NN). The output of the network is a vector of dimension 2, values corresponding to the probability that the data tuple belongs to particular class. Type Input Flatten Maps and Neurons 16M × 64 × 7N 7168N Filter Size Table S3: Network architecture used for subject discrimination network (adversary). The output of the network is a vector of dimension 25, values corresponding to the probability that the data tuple belongs to particular subject. 1.2 Topographical maps Layer 0 1 2 3 4 5 6 7 8 9 Type Input Conv ReLU Conv ReLU MaxPool MaxPool Conv ReLU MaxPool Maps and Neurons 3M × 64 × 64N 16M × 32 × 32N 16M × 32 × 32N 16M × 16 × 16N 32M × 16 × 16N 32M × 16 × 16N 32M × 8 × 8N 64M × 8 × 8N 64M × 8 × 8N 64M × 4 × 4N Filter Size - 5×5 - 2×2 5×5 - 2×2 3×3 - 2×2 Table S4: Network architecture used for EEG feature extraction network (Base CNN). The output of the network is a tensor of dimensions 64 × 4 × 4. Type Input Flatten Maps and Neurons 64M × 4 × 4N 1024N Filter Size Layer 0 1 2 3 4 5 Layer 0 1 2 3 4 5 Dropout (p=0.5) FullyConn ReLU FullyConn Dropout (p=0.5) FullyConn ReLU FullyConn - 8N 8N 2N - 8N 8N 25N - - - 1×1 - 1×1 - - - 1×1 - 1×1 Table S5: Network architecture used for group discrimination network (Top NN). The output of the network is a vector of dimension 2, values corresponding to the probability that the data tuple belongs to particular class. Type Input Flatten Maps and Neurons 64M × 4 × 4N 1024N Filter Size Table S6: Network architecture used for subject discrimination network (adversary). The output of the network is a vector of dimension 25, values corresponding to the probability that the data tuple belongs to particular subject. 3 2 Training curves 2.1 Time-Frequency Maps Hyperparameter Learning Rate Learning Rate Decay Weight Decay Value 0.001 0.0001 0.001 Table S7: List of hyperparameters used for training the networks on TF maps. (a) Group prediction loss (λ = 0) (b) KL divergence (λ = 0) (c) Subject prediction loss (λ = 0) (d) Group prediction loss (λ = 15) (e) KL divergence (λ = 15) (f) Subject prediction loss (λ = 15) (g) Group prediction loss (λ = 13) (h) KL divergence (λ = 13) (i) Subject prediction loss (λ = 13) Figure S2: Time-Frequency Maps Training curves for three different weight values to the subject predictor regularizer. 4 2.2 Topographical Maps Hyperparameter Learning Rate Learning Rate Decay Weight Decay Value 0.001 0.001 0.03 Table S8: List of hyperparameters used for training the networks on Topographical maps. (a) Group prediction loss (λ = 0) (b) KL divergence (λ = 0) (c) Subject prediction loss (λ = 0) (d) Group prediction loss (λ = 10) (e) KL divergence (λ = 10) (f) Subject prediction loss (λ = 10) (g) Group prediction loss (λ = 5) (h) KL divergence (λ = 5) (i) Subject prediction loss (λ = 5) Figure S3: Topographical Maps Training curves for three different weight values to the subject predictor regularizer. 5 3 ccCAM on Topographical Maps Figure S4: ccCAM matrices generated for Topographical Maps of 23–33 Hz activity group-averaged over time windows of 0.6 sec. The CAMs have higher values for EXE group as compared to the CON group which indicates that the network learns to look at the differences in EEG activity while performing the fixed-force handgrip task induced by acute bout of high-intensity exercise. The time described here is calculated from 0.5 sec after the presentation of visual cue. 6
1509.07897
1
1509
2015-09-16T06:02:26
Quantum Look at two Common Logics: the Logic of Primitive Thinking and the Logic of Everyday Human Reasoning
[ "q-bio.NC", "cs.AI", "math.LO" ]
Based on ideas of quantum theory of open systems and psychological dual system theory we propose two novel versions of Non-Boolean logic. The first version can be interpreted in our opinion as simplified description of primitive (mythological) thinking and the second one as the toy model of everyday human reasoning in which aside from logical deduction, heuristic elements and beliefs also play the considerable role. Several arguments in favor of the interpretations proposed are adduced and discussed in the paper as well.
q-bio.NC
q-bio
Quantum Look at two Common Logics: the Logic of Primitive Thinking and the Logic of Everyday Human Reasoning B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Sciences of Ukraine 47, Lenin Ave., Kharkov 61103, Ukraine. (Dated: October 24, 2018) E. D. Vol∗ Based on ideas of quantum theory of open systems and psychological dual system theory we propose two novel versions of Non-Boolean logic. The first version can be interpreted in our opinion as simplified description of primitive (mythological) thinking and the second one as the toy model of everyday human reasoning in which aside from logical deduction, heuristic elements and beliefs also play the considerable role. Several arguments in favor of the interpretations proposed are adduced and discussed in the paper as well. PACS numbers: 05.40.-a I. INTRODUCTION It is well known that the Boolean propositional logic which is consistent mathematical presentation of the clas- sical Aristotelian logic has applications of two kinds. On the one hand it correctly takes into account deductive elements of the everyday human reasoning and on the other hand this logic can be considered as the relevant framework of common scientific language both in exact sciences and humanities. For example referring to the classical physics it may be argued that the result of any experiment realized in this area can be described by the Boolean logic in a consistent manner. However in the sit- uations when distinct features of the objects under study strongly correlate between themselves and moreover may be incompatible the application of standard Boolean logic may lead to errors. Just such case has place for example in quantum mechanics. Actually, as firstly G. Birkhoff and J. von Neumann revealed [1] in quantum mechanics, in view of existence of certain non-commuting observ- ables relating to the same particle such as coordinate and momentum or the different components of the spin, the laws of the Boolean logic (in particular the distributive law) have been failed and their generalization i.e. quan- tum logic is desirable. The similar situation may arise also in everyday life when various reasons, emotions and beliefs governing the behavior of concrete person begin to contradict each other. In this connection it should be noted that still in the early days of quantum theory N. Bohr, when he formulated the Complementarity Prin- ciple, (which just maintains the presence in physics cer- tain supplementary properties of the microscopic objects) underlined more than once essential analogy existing be- tween atomic processes and such mental phenomena as thoughts, sentiments and acts of decision making. It is remarkable that modern cognitive psychology using the data of a large number of experiments came to the similar conclusions as well. Really according to the psychologi- ∗Electronic address: [email protected] cal dual system theory all basic cognitive processes such as attention, memory, learning and so on are connected with certain dual systems and dual mechanisms. In the present paper we will interested in only human reasoning mechanisms where two primary dual systems of interest can be specified. The first of them (we will term it below as deductive reasoning system) is rational, sequential and consistent but acts relatively slow and its resources are limited by the capacity of human working memory. The second one (we will term it as heuristic reasoning system ) is intuitive, rapid and automatic but its activity may be biased to a large degree by emotions and old unconscious ideas. In addition a hidden interaction exists between these two cognitive systems which as a rule is not aware by the reasoning person. The main goal of present pa- per is to demonstrate how by using some ideas, concepts and technical tools of quantum theory of open systems (QTOS) one can describe the dual nature of human rea- soning in the framework of consistent logical theory that generalizes and modifies the rules of ordinary Boolean logic. The paper is organized as follows. In Section.2 fol- lowing the previous author paper [2] we briefly outline main ideas of the approach proposed using the simple and instructive example of probabilistic Boolean logic (PBL) that describes the situations when in the absence of complete information about the surrounding objects and events all human judgments acquire inevitably prob- abilistic nature. To take into account this crucial point it is convenient, instead of usual Boolean functions of dis- crete variables which take only two values unit and zero, to associate with every plausible proposition (PP) certain diagonal representative density matrix (RDM) of some auxiliary two state quantum system, whose elements de- fine the plausibility of corresponding proposition. The essential and novel element of the approach proposed is the universal and constructive algorithm (based on QTOS) which allows one to define all logical connec- tives between plausible propositions using the powerful and effective method of positive valued (PV) quantum operations. It is worth noting that this method will be used continually throughout the paper for the construc- tion of various kinds of logics. Also we study here an- other important problem, namely : how the possible log- ical correlations between various PP may be took into account in the approach proposed. Further in the Sec- tion.3 in order to describe the dual nature of human rea- soning we have extended our approach to the arbitrary 2 × 2 non-diagonal density matrices that will represent in this case the generalized propositions (GP). We as- sume that non-diagonal elements of the RDMs of such propositions relate to heuristic ( believable) elements of human reasoning and therefore by means of them one can define the believability of the corresponding propo- sitions. The main problem of the approach proposed is: whether one may specify the set of logical connectives for similar generalized propositions and if the answer is yes in what way. To answer this question we will use again the approved method of PV quantum operations. It turns out however that in general case (beyond the usual PBL) it is necessary to impose certain restrictions either on the form of admissible quantum operations or on the form of input GP. In this connection we proposed two possible alternatives to construct the consistent Non- Boolean (NB) logic operating with such GP. The NB logic of the first kind aside from negation includes only the single pair of two place connectives which are simi- lar to the pair of relations: equivalence-non-equivalence in the standard Boolean logic. At first sight such logic looks much poorer than standard Boolean logic and we will term it as 'atomic' logic or prime logic. It turns out however that, starting from GP and using only these two connectives, one may construct any prescribed diagonal PP and after that to handle with these PP according to the known rules of PBL .Thus PBL appears now as secondary or 'molecular' logic with respect to prime or 'atomic' NB logicof the first kind. We argue that described version of Non-Boolean logic of the first kind (prime logic) can be considered as sim- plified version of the logic of primitive (or mythological) thinking. In favor of such interpretation we adduce sev- eral cogent in our opinion arguments. Finally in the Sec- tion.4. we study the NB logic of the second kind that assumes the existence of special primordial correlations between deductive and heuristic components of general- ized propositions .In this case using the formalism of PV quantum operations one can specify all the same logical connectives between GP including implication as in usual PBL .It turns out that a number of known from everyday life curious psychological phenomena such as belief-bias effect may be explained in the framework of this logic by natural way. Now let us go to the concrete presentation of announced results. II. PRELIMINARIES In this section we briefly remind some results of our previous paper [2] in which based on ideas of quantum theory of open systems the formalism of probabilistic 2 Boolean logic (PBL) was developed. To this end in view, instead of standard Boolean variables with two values 1 and 0 that represent true and false propositions respec- tively one has to consider so called plausible propositions (PP), whose truth or falsity are determined only with certain probability. We propose to represent such propo- sitions by means of 2 × 2 diagonal matrices with positive elements the sum of which is equal to unit. Every such representative matrix of PP may be associated with the density matrix of certain two state quantum system . Therefore, if it does not lead to confusion, we will often identify the propositions in PBL with their representative density matrices (RDMs). In explicit form the RDM of arbitrary PP a looks as 0 1 − pa(cid:19) where pa is a probability for propo- ρ (a) =(cid:18)pa 0 sition a to be true(index a of the proposition we will often omit later ). It is convenient also to define the plausibility of the proposition a as P (a) = 2pa − 1. Ob- viously that plausibility takes its values in the interval [−1, 1]. It turns out the approach proposed allows one to express all logical connectives between PP by universal way as certain positive valued (PV) quantum operations under their RDMs. Referring the reader for the details of this approach to [2] let us merely demonstrate here how this approach works using several concrete examples. So, if someone wants to obtain the RDM of the negation of PP a =(cid:18)p with the following 2×2 matrix Gnot =(cid:18)0 1 0 1 − p(cid:19) that is (not a) she(or he) must make 1 0(cid:19) and GT is, the next quantum operation: ρ (nota) = Gnotρ (a) GT 0 not as usually, the matrix transposed to matrix G. It turns out that any two place connective in PBL (aRb) may be expressed in similar manner as well, namely: (aRb) = GR [ρ (a) ⊗ ρ (b)] GT R (1) where ρ (a) ⊗ ρ (b) is standard tensor product of RDMs of propositions a and b and GR is a 2 × 4 matrix (we will term it below as admissible matrix for connective R) which has two defining properties:1) every element of GR is equal 0 or 1 and 2) in each column of matrix GR the only element is equal to one and all the rest are equal to zero. Let us now present (without proof) the concrete form of matrices GR for basic two place connectives, namely :(a and b), (a or b),(a =⇒ b) . As shown in [2] they have the next form: 0 1 1 1(cid:19) , G(a and b) = (cid:18)1 0 0 0 0 0 0 1(cid:19) , G(a or b) = (cid:18)1 1 1 0 0 1 0 0(cid:19) . G(a=⇒b) = (cid:18)1 0 1 1 (2) (3) (4) For example for the implication (a =⇒ b) one can easily 3 (5) 1) ρ1 (a, b) =     1 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 2 0 0 0 0 0 1 2 0 0 0 0 1 2 0 0 0 0 0  ,   and 2) ρ2 (a, b) = 3) ρ3 =   1 4 0 0 0 0 1 4 0 0 0 0 1 4 0 0 0 0 1 4  . writes down the RDM as: 0 ρ (a =⇒ b) =(cid:18)1 − p + pq p (1 − q)(cid:19) . In Eq. (5) we mean the next notation: a = (cid:18)p and b = (cid:18)q 0 1 − p(cid:19), 0 1 − q(cid:19). Note that this relation coincides 0 0 0 with standard Boolean expression for the implication in the case when probabilities p and q may take only two values 0 and 1. Let us make now the generalization of the approach proposed on the important case when ini- tial PPs a and b are not independent propositions or, in other words, they possess some logical correlations. In such situation the best way to take these correla- tions into account is to avail of the direct analogy be- tween our approach and the theory of composite cor- related quantum systems. To realize this analogy one must take as a starting point, instead of tensor produc- tion of two initial PPs a and b the more general positive diagonal matrix of the form: ρ (a, b) =   4Pi=1 with additional normalization condition: p1 0 0 0 p2 0 0 0 0 0 0 p3 0 0 0 p4   pi = 1. Fur- thermore we naturally assume that this joint matrix of correlated propositions ρ (a, b) is connected with matri- ces of partial propositions that is RDM of a and RDM of b by the standard relations: a = (cid:18)p1 + p2 b = (cid:18)p1 + p3 p3 + p4(cid:19), p2 + p4(cid:19). Now let us suppose again that 0 0 0 0 all logical connectives for correlated PP may be obtained by the approved method that is by means of the PV quan- tum operations. Here we restrict ourselves only with the case of implication between correlated propositions a and b. Let us introduce the following notation p = p1 + p2 , q = p1 + p3 and C = p1p4 − p2p3 0 pq + C p (1 − q) − C It is easy to see that ,using this notation, the ma- trix ρ (a, b) can be rewritten in the form: ρ (a, b) =   0 1 0 0(cid:19) to the ρ (a, b) one may obtain the required (cid:18)1 0 1 1 (1 − p) (1 − q) + C Now by applying the quantum operation G(a=⇒b) = q (1 − p) − C 0 0 0 0 0 0 0 0 0 0 0  . result for RDM of implication, namely: ρ (a, b) =(cid:18)1 − p + pq + C 0 p − pq − C(cid:19) . 0 (6) It is easy to see that in all cases the RDMs for partial propositions a and b coincide and represent the same ambiguous proposition: ρi (a) = ρi (b) = 2(cid:19) (i = 1, 2, 3). However the expressions for impli- (cid:18) 1 2 0 0 1 cation in these situations differ significantly. Really us- ing the general rule Eq. (1) stated above and its spe- cial case in the form of Eq. (6) one may easily obtain: (a =⇒ b)1 = (cid:18)1 0 (a =⇒ b)2 =(cid:18) 1 and (a =⇒ b)3 = (cid:18) 3 0 0(cid:19), that is the true proposition, while 2(cid:19) is the same ambiguous proposition 4(cid:19) .Note that only the last case 4 0 0 1 2 0 0 1 with composite matrix ρ3 corresponds to the case of the independent partial propositions. This simple but in- structive example clearly demonstrates the considerable influence of logical correlations on the results of inference process. In this point we have finished our brief review of PBL. It should be noted in conclusion that in the case of PBL considered in this Section were taken into ac- count only propositions that may be specified by the sin- gle defining characteristic , namely plausibility. Now we pass to the study of Non-Boolean (NB) logics that would operate with the propositions of more general kind .These propositions can be determined already by two defining features which reflect dual nature of the objects (or sub- jects) connecting with corresponding propositions . We believe that similar logics may be more appropriate for the description (at least in simplified form) of essential peculiarities of such common logics as the logic of prim- itive (mythological) thinking and the logic of everyday human reasoning. III. NON-BOOLEAN LOGIC OF THE FIRST KIND AS THE SIMPLIFIED DESCRIPTION OF PRIMITIVE (MYTHOLOGICAL) THINKING Note that the variable C is just a measure of logical corre- lations between propositions a and b. In this connection we will termed it as the context variable. Let us consider now the simple example with three con- crete initial RDMs of composite correlated propositions, namely : In the beginning of this Section we would like to dis- cuss the following primary question namely: to what de- gree the approach stated above may be extended on the propositions of more general form .To this end in view let us consider instead of exclusively diagonal RDM rep- resenting a certain PP the more general class of 2 × 2 non-diagonal positive valued matrices with unit trace and examine them as certain RDMs of generalized propo- sitions (GP). Let us suppose that an arbitrary propo- sition of this class a can be written in the next form z ∗ 1 − p(cid:19) ,where z is some complex number. ρ (a) = (cid:18) p p(cid:19). is (not a) by natural way as ρ (not a) = (cid:18)1 − p z ∗ Let us define also the negation of such proposition, that z z Furthermore let us require that the RDMs of the oppo- site propositions a and not a commute with each other. The sense of this restriction (having in mind the obvious analogy with quantum theory) is quite clear and does not need in additional comments. It is easy to see that this restriction may be reduced to the simple relation z = iα or, in other words, non-diagonal elements of the RDM in the approach proposed have to be pure imaginary. Thus the class of generalized RDMs which we will consider in the present Section contains all 2 × 2 positive valued ma- trices of the form: ρ (a) =(cid:18) p −iα 1 − p(cid:19) . iα 4 we have to define all possible connectives related to them. It turns out however that the direct extension of the ad- missible quantum operation method in order to obtain required two place logical operations for GPs is impossi- ble ( because the automatic satisfaction of normalization condition, that was guaranteed for PPs in PBL, is not always holds now).However there are at least two possi- ble ways to get round this formal obstacle. In this Sec- tion we consider the first of them. To this end we retain as admissible operations only two concrete 2 × 4 matri- ces,namely: G∆ = (cid:18)0 1 1 0 0 1 1 0(cid:19). 1 0 0 1(cid:19) and G∆ = (cid:18)1 0 0 1 It is easy to see directly that these operations transform the tensor product of two propositions a ⊗ b to the RDM that belongs to the required class of GP . Really, let a = (cid:18) p −iα 1 − p(cid:19) and b = (cid:18) q −iβ 1 − q(cid:19) are two GP, iα iβ then it is easy to obtain that: G∆ (a ⊗ b) GT ∆ =(cid:18) A −iγ 1 − A(cid:19) ≡ (a∆b) . iγ (9) (7) where A = p + q − 2pq + 2αβ and γ = α (1 − 2q) + β (1 − 2p). In the same way one can find that As to formal interpretation of these GPs we assume as before that their diagonal elements determine the plausi- bility of corresponding GP according to the above men- tioned rule: P (a) = 2p − 1 whereas their non-diagonal elements define its believability, that is the readiness of a subject to accept the corresponding proposition as valid by virtue of some irrational and unconscious reasons or, in other words. on the strength of certain beliefs. Thus the similar GPs combine together the objective and sub- jective reasons which stimulate a person to accept the concrete proposition as valid. It is naturally to define the believability of concrete GP a with RDM Eq. (9) as B (a) = −2α. The justification of such definition became clear if we write down this GP by standard way using the Bloch sphere representation, namely: G∆ (a ⊗ b) GT ∆ = (cid:18)1 − A −iγ A (cid:19) = iγ (10) = not(a∆b) ≡ (a∆b). In the case when x-component of the Bloch vector is identically equal to zero it is convenient to intro- duce the complex Bloch vector as follows : P = Pz − iPy.Comparing this definition with the above mentioned Bloch sphere representation of density matrix one can state the simple relation connecting the complex vector R of the proposition (a∆b)with corresponding vectors P and Q of the GP a and b respectively, namely: R = −P Q. (11) ρ (a) =(cid:18) 1+Pz Px+iPy 2 2 Px −iPy 2 2 (cid:19) . 1−Pz (8) Note that for GPs which we consider here, the compo- −→ nent Px of the Bloch vector P is always equal to zero. Now if one compares the representation Eq. (8) with ex- pression Eq. (7) she(he) can see that the plausibility and believability of the same GP being expressed in terms of the Bloch vector components look similarly. Furthemore it should be noted that relation P 2 y 6 1 is obviously holds for any GP and this fact distinctly reflects the com- petition existing between two features of the same propo- sition. It should be noted also that our assumption is in a full agreement with modern psychological dual system theory [3] that asserts the presence in brain two com- plementary cognitive systems: (deductive and heuristic) , which both are responsible for human reasoning, al- though often compete with each other. Now in order to formulate the consistent NB logic with GPs stated above z + P 2 So, we found that the NB logic of the first kind aside from negation contains in addition only the single pair of two place connectives, namely ∆ and ∆. At first sight such logic is much poorer that standard Boolean logic and therefore one may call it as 'atomic logic' or prime logic. Nevertheless it should be noted that, having in hands only these two connectives and starting from the set of GPs, one can easily obtain any PP belonging to PBL with prescribed diagonal elements. After making that, one may operate with them already according to the rules of standard Boolean logic. Let us demonstrate the validity of this statement .To this end let us take two opposite GP of the special form: a = (cid:18) 1 (not a) =(cid:18) 1 2(cid:19) and 2 (cid:19). If one applies the operation ∆ to 2 −iα iα iα −iα 1 2 1 them the obtained result looks as: 2 − 2α2 (a∆ not a) =(cid:18) 1 0 0 2 + 2α2(cid:19) . 1 (12) 2 Thus we see that any proposition a of PBL with pa 6 1 2 may be expressed by similar way. As for propositions of PBL with pa ≥ 1 2 they can be expressed by means of op- eration ∆ in the same manner. Thus the PBL plays now the role of secondary or 'molecular' logic with respect to prime or 'atomic' NB logic of the first kind. Note in addition that the GP of the form : a = (cid:18) 1 iα −iα 1 2(cid:19) whose plausibility is equal to zero can be considered as the simplest example of so-called intuitive judgments. In this place we refer the reader to the inspiring book [4] in which distinction between discoursive and intuitive judg- ments have been deeply analyzed. Now let us pass to the most intriguing question, relating to the NB logic of the first kind stated above, namely: whether the logic proposed above has any actual applications in everyday life and (or) in science? We believe that the answer is yes and are going to argue that logic of this kind could be appropriate for example as simplified description of primitive ( mythological) thinking. The author is aware that without being an expert in the area of anthropology or mythology he is unable to disclose this topic in proper depth and moreover to prove the above assertion exactly. Nevertheless let me adduce several arguments in favor of the hypothesis proposed above. In the beginning let us- remind some basic facts relating to the primitive thinking (PT), or the thinking of savages ( all these and the other concrete facts the reader can derive in detail for exam- ple from classical book by C. Levi-Strauss) [5].The major fact on which we base is that there are a set of elemen- tary and at the same time fundamental units of PT, so- called binary oppositions, such as : top-bottom, right- left, birth-death, male-female and so on. Another impor- tant regularity determines the basic principle governing the primitive thinking, that is, its goals and mental tools which are used to achieve these goals . It turns out that PT while it operates with binary oppositions does not seek to avoid any contradictions, that is typical for the "normal" logic, but rather tends to find and then to re- construct all kinds of possible intermediate links existing between two antagonistic terms in concrete binary oppo- sition. To this end in view there are two special person- ages in myth, namely: the Culture Hero and the Trixter, (for example in Greek mythology Hermes plays the role of such Trixter) which realize this important function. Fur- thermore certain totem animals for example the Raven and the Coyote in Indian myths sometimes play the part of Culture hero as well. The basic tool using in PT that allows one to achieve the desired goal is so-called 'medi- ation' -- the special mental or workable operation which is able to reduce the distinction between two antagonis- tic terms in binary oppositions. For example the Raven and the Coyote as the animals- scavengers symbolically realize the mediation function between the world of the living beings and the world of the dead . Now let us demonstrate how in the NB logic of the first kind de- scribed above its two basic operations may be considered as certain analogues of the mediation operation in myth. 5 To this end, in order to define the degree of difference be- tween two distinct GPs, we will use well-known definition of the distance between two mixed quantum states ρ1 and ρ2, namely D (ρ1, ρ2) = 1 2 Sp ρ1 − ρ2 [6]. It is easy to see that in the case of two dimensional Hilbert space and in the Bloch sphere representation the above expres- sion takes the simple form: D (ρ1, ρ2) = 1 where the states. −→ P 2 are corresponding Bloch vectors of P 2(cid:12)(cid:12)(cid:12) −→ P 1 and 2(cid:12)(cid:12)(cid:12) −→ P 1 − −→ It is obvious that maximum distance is realized be- tween opposite propositions or , (using the PT language) between antagonistic members of binary opposition. Now if one applies the basic two place operation ∆ to the pair of opposite propositions a with complex vector P and (nota) with complex vector −P she(he) obtains the new proposition with the complex vector R = P 2. It is easy to see that the distance between the opposite proposi- tions with complex vectors P 2 and −P 2 is less that the distance between original pair of opposite propositions. This fact confirms that logical connective ∆ possesses the characteristic property of operation 'mediation' in PT. Now we are going to adduce another forcible argument in favor of the doubtless connection between NB logic of the first kind considered above and the logic of PT. Let us remind that according to the known ethnologist Levy-Bruhl [7] (who was the first European researcher of primitive and magical thinking PT is governed to a great extent by the so-called law of participation that claims the universal links existing between various things and events in the world.In particular just this law allows the shaman, who fell into trans, to perform a wide variety of magical acts and transformations with surrounding ob- jects or people using their mutual contiguity and (or) similarity. It is interesting to note that this feature of primitive thinking afterwards was adopted by art and literature (especially poetry) in the form of extensive use of such specific tropes as metaphor and metonymy. In the NB logic of the first kind proposed above this fea- ture of PT may be simulated by additional logic opera- tions,namely by logic rotations, that have no analogues in the standard Boolean logic. Really, it is possible to define one-parameter continuous group of logic rotations (in the plane Pz − Py) according to the usual mathemati- cal rule, namely: let GP a is represented by the complex vector P then GPea obtained by rotation of a at an angle Φ has complex vector eP with components fPy and fPz so thatfPy = Py cos Φ+Pz sin Φ andfPz = Pz cos Φ−Py sin Φ .Clearly that in this groupof logical transformations the negation of the proposition a coincides with its rotation at angle π and, furthermore, if one rotates GP a at angle Φ1 and the other GP b at angle Φ2 then the GP (a∆b) would be rotated at angle (Φ1 +Φ2).Thus this continuous group of logical rotations let one to transform any propo- sition into the arbitrary other (with the same modulus certainly). In this point we are obliged to finish our brief discussion of the relationship existing between primitive thinking and the NB logic of the first kind although it is clear that this complex and intriguing topic deserves much more detail study. IV. NON-BOOLEAN LOGIC OF THE SECOND KIND AS THE TOY MODEL OF EVERYDAY HUMAN REASONING As we see in the previous Section in the case of Non- Boolean logic of the first kind it is possible to define aside from negation only the single pair of two- place logical op- erations ∆ and ∆. The application of the rest admissible quantum operations to the tensor product of two input GP results in the disturbance of normalization condition for output RDM. This obstacle may be formally removed if one assume the existence of certain primordial correla- tions between plausible and believable components of the GPs. Let us explain more detail what we have in mind. Suppose that instead of tensor product of the two input GP a⊗ b (that means their logical independence) we take as input more general composite proposition ρ (a, b) , the RDM of which has the next form: ρ (a, b) =  p1 −iα p2 0 −iβ iα iβ 0 p3 0 iγ iδ 0 −iγ −iδ p4   . (13) Using the obvious analogy with theory of compos- ite quantum systems one can determine now the partial RDM of propositions a and b as a = −i (β + γ) p3 + p4 (cid:19) and b = (cid:18) p1 + p3 (cid:18) p1 + p2 −i (α + δ) p2 + p4(cid:19) i (β + γ) i (α + δ) respectively. It is clear that input composite proposition ρ (a, b) Eq. (13) corresponds to the case of two corre- lated partial GP a and b and the degree and nature of their correlation is uniquely determined by elements of composite matrix ρ (a, b). It should be emphasized that construction of the input composite proposition ρ (a, b) in the form Eq. (13) has the essential advantage over simple tensor product from formal point of view since it allows one to define all 16 two place logical connec- tives consistently by approved method of PV quantum operations. Significantly that all admissible matrices GR in this NB logic of the second kind have the same form as in the case of PBL (see Eq. (1) ) and corresponding two place connectives can be obtained according to the similar rule: (aRb) = GRρ (a, b) GT R. (14) Let us present here (without details) the expressions for the two place connectives (and), (or) in the NB logic of the second kind, obtained by means of Eq. (14): p1 (a and b) = (cid:18) −i (α + β) p2 + p3 + p4(cid:19) , (a or b) = (cid:18)p1 + p2 + p3 i (γ + δ) p4 (cid:19) −i (γ + δ) i (α + β) (15) (16) 6 Now as in the previous section, the important question arises about the possibility of concrete realizations of the NB logic of the second kind. In this connection we are going to bring several argu- ments in favor of hypothesis that above version of NB logic can be considered as simplified model of everyday human reasoning. To this end in view let us remind some basic facts relating to the psychological dual system the- ory of human cognition(see for example [3] for the com- prehensive exposition). As numerous psychological ex- periments and observations had unambiguously demon- strated there are two distinct types of mental processes that are responsible for human reasoning: 1) cognitive processes of the first type that are rapid, automatic and intuitive and 2) the processes of the second type that are slow, sequent and rational. Cognitive system which is responsible for the processes of the first type is termed as heuristic (or intuitive) while the second one is termed as deductive (or analytical). Clearly these two systems may sometimes come into conflict and compete with each other. Therefore human reasoning is often biased and subjected to various fallacies. In particular the reason- ing person often overestimates the cogence of the argu- ments that lead to the believable and expected conclu- sions and underestimates those arguments which lead to the conclusions contradictory to her (his) beliefs. This widespread psychological phenomenon is known as belief- bias effect. Now the natural question arises: is it possible to describe this human hybrid intuitive-deductive think- ing in the framework of any consistent logical theory? In this Section we propose brief outline of possible the- ory based on the NB logic of the second kind that was exposed above. Note that we present here only the lit- tle fragment of this theory which let one to understand belief-bias effect in the framework of closed logical point of view. To this end let us examine the expression for implication connecting two correlated GP. According to the general rule Eq. (14) the RDM of the implication can be represented in the form: (a =⇒ b) = Gimpρ (a, b) GT imp =(cid:18)p1 + p3 + p4 i (α − γ) p2 (cid:19) . 0 1 0 0(cid:19) . where the matrix Gimp =(cid:18)1 0 1 1 −i (α − γ) (17) We see that the believability of the GP (a =⇒ b) as it follows from Eq. (17) is equal to 2 (γ − α), and it does not connect by simple way with corresponding believabil- ities of partial propositions a and b. Let us now impose another additional restrictionon on the phases α, β, γ, δ of the matrix ρ (a, b) which looks enough naturally, namely, we require that in simple version of NB logic of the sec- ond kind the known De Morgan dual formulas for all logical propositions, namely: not (a and b) = (not a) or (not b) and not (a or b) = (not a) and (not b) hold. This is just the case what we have in mind when we are speaking about the toy model of everyday human reasoning. Really, it is easy to see that the restriction above immediately leads to the next condition imposed on the phases of input matrix ρ (a, b) of Eq. (13), namely: α+β = γ+δ = const. Since this constant must be identi- cal for all propositions in the toy model of NB logic of the second kind we can without loss of generality set it equal to zero and then it turns out that all non-diagonal ele- ments in implication ( and in other connectives as well) can be expressed by means of the non-diagonal elements of partial propositions. Comparing the believability of the implication (a =⇒ b) with the believability of it con- sequence b that is equal to −2 (α + δ) = 2 (γ − α)we im- mediately come to the desired conclusion , namely, in the toy version of NB logic of the second kind the values of believabilities of the propositions (a =⇒ b) and b com- pletely coincide (and both are equal to 2 (γ − α)).Strictly speaking just this conclusion gives us the good reason to believe that NB logic of the second kind has essential common features with everyday human logic in which (as was shown in numerous experiments) the believabil- ity of conclusions stimulates the subjects more positively evaluate the correctness of the process of their inference and to pay less attention to their logical rigor (belief-bias effect). It is necessary to mention also that as a matter of fact the whole subjective evaluation of the validity of 7 the given proposition a simultaneously should take into account both the plausibility of this proposition and its believability as well. Therefore the appropriate expres- sion for the integral validity V of the GP a should be actually look as follows: V (a) = αP (a) + (1 − α) B (a) . (18) (where the coefficient α reflects the mental nature of rea- soning person.It remains to point out that by appli- cation of the results of the approach proposed in this Section one can evaluate quantitatively both believabil- ity and plausibility of various GPs in complex logical situations relating to everyday human reasoning. How- ever this amusing task is already beyond the scope of the present paper. In the conclusion we should like to emphasize only that all guesses and conjectures about proposed interpreta- tions of two NB logics expressed in this paper must be taken with a grain of salt, that is as plausible but prelim- inary hypotheses . Obviously, the rigorous confirmation of all results obtained here will require more painstaking extra work in the future. [1] G. Birkhoff, J. von Neumann, Ann.Math.,37,823, (1936) [2] E.D. Vol, Int.J. of Theor. Phys.,52,514,(2013) [3] Jonathan Evans,Keith Frankish, In Two Minds: Dual Pro- [6] M.A. Nielsen, I.L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press, (2001) cesses and Beyond,Oxford University Press,(2009 ) [7] L. Levy-Bruhl, La Mythologie Primitive, Paris, (1935) [4] E.L. Feinberg, Zwei Kulturen, Springer-Verlag,(1998) [5] C. Levi-Strauss, La Pensee Sauvage, Paris, (1962)
1111.6573
1
1111
2011-11-28T20:23:29
Neural integrators for decision making: A favorable tradeoff between robustness and sensitivity
[ "q-bio.NC" ]
A key step in many perceptual decision tasks is the integration of sensory inputs over time, but fundamental questions remain about how this is accomplished in neural circuits. One possibility is to balance decay modes of membranes and synapses with recurrent excitation. To allow integration over long timescales, however, this balance must be precise; this is known as the fine tuning problem. The need for fine tuning can be overcome via a ratchet-like mechanism, in which momentary inputs must be above a preset limit to be registered by the circuit. The degree of this ratcheting embodies a tradeoff between sensitivity to the input stream and robustness against parameter mistuning. The goal of our study is to analyze the consequences of this tradeoff for decision making performance. For concreteness, we focus on the well-studied random dot motion discrimination task. For stimulus parameters constrained by experimental data, we find that loss of sensitivity to inputs has surprisingly little cost for decision performance. This leads robust integrators to performance gains when feedback becomes mistuned. Moreover, we find that substantially robust and mistuned integrator models remain consistent with chronometric and accuracy functions found in experiments. We explain our findings via sequential analysis of the momentary and integrated signals, and discuss their implication: robust integrators may be surprisingly well-suited to subserve the basic function of evidence integration in many cognitive tasks.
q-bio.NC
q-bio
Title: Neural integrators for decision making: A favorable tradeoff between robustness and sensitivity Abbreviated Title: Neural integrators for decision making Authors: Nicholas Cain, Department of Applied Mathematics, University of Washington Andrea K. Barreiro, Department of Applied Mathematics, University of Washington Michael Shadlen, Department of Physiology and Biophysics, University of Washington Eric Shea-Brown, Department of Applied Mathematics, University of Washington Corresponding Author: Eric Shea-Brown Department of Applied Mathematics Box 325420 Seattle, WA 98195-2420 [email protected] Number of pages: 30 Number of figures: 16 Number of tables: 1 Number of words (Abstract): 226 Number of words (Introduction): 476 Number of words (Discussion): 1467 Conflict of Interest: None Acknowledgements: This research was supported by a Career Award at the Scientific Interface from the Burroughs-Wellcome Fund (ESB), the Howard Hughes Medical Institute, the National Eye Institute Grant EY11378, and the National Center for Research Resources Grant RR00166 (MS), by a seed grant from the Northwest Center for Neural Engineering (ESB and MS), and by NSF Teragrid allocation TG-IBN090004. arXiv:1111.6573v1 [q-bio.NC] 28 Nov 2011 1 Abstract A key step in many perceptual decision tasks is the integration of sensory inputs over time, but fundamental questions remain about how this is accomplished in neural circuits. One possibility is to balance decay modes of membranes and synapses with recurrent excitation. To allow integration over long timescales, however, this balance must be precise; this is known as the fine tuning problem. The need for fine tuning can be overcome via a ratchet-like mechanism, in which momentary inputs must be above a preset limit to be registered by the circuit. The degree of this ratcheting embodies a tradeoff between sensitivity to the input stream and robustness against parameter mistuning. The goal of our study is to analyze the consequences of this tradeoff for decision making performance. For concreteness, we focus on the well-studied random dot motion discrimination task. For stimulus parameters constrained by experimental data, we find that loss of sensitivity to inputs has surprisingly little cost for decision performance. This leads robust integrators to performance gains when feedback becomes mistuned. Moreover, we find that substantially robust and mistuned integrator models remain consistent with chronometric and accuracy func- tions found in experiments. We explain our findings via sequential analysis of the momentary and integrated signals, and discuss their implication: robust integrators may be surprisingly well-suited to subserve the basic function of evidence integration in many cognitive tasks. 1 Introduction Many decisions are based on the balance of evidence that arrives at different points in time. This process is quantified via simple perceptual discrimination tasks, in which the momentary value of a sensory signal carries negligible evidence but correct responses arise from summation of this signal over the duration of a trial. At the core of such decision making must lie neural mechanisms that integrate signals over time (Gold and Shadlen, 2007; Wang, 2008; Bogacz et al., 2006). The function of these circuits is intriguing, because perceptual decisions develop over hundreds of milliseconds to seconds, while individual neuronal and synaptic activity often decays on timescales of several to tens of milliseconds -- a difference of at least an order of magnitude. A mechanism that bridges this gap is feedback connectivity tuned to balance -- and hence cancel -- inherent voltage leak and synaptic decay (Cannon et al., 1983; Usher and McClelland, 2001). The tuning of recurrent connections to achieve this balance presents a challenge (Seung, 1996; Seung et al., 2000), illustrated in Figure 4(A) via motion of a ball on an energy surface. Here, the ball position E(t) represents the total activity of a circuit (relative to a baseline marked 0); momentary sensory input perturbs E(t) to increase or decrease. If decay dominates (upper-right), then E(t) always has a tendency to "roll back" to baseline values, thus forgetting accumulated sensory input. Conversely, if feedback connections are in excess, then activity will grow away from the baseline value (center). If balance is perfectly achieved via fine-tuning, (left) temporal integration can occur. That is, inputs can then smoothly perturb network activity back and forth, so that the network state at any given time represents the time-integral of past inputs. Koulakov et al. (2002) proposed an alternate model: a ratchet-like accumulator, equivalent to movement along a scalloped energy surface (Figure 4(A), bottom) (Pouget and Latham, 2002). Importantly, even without finely-tuned connectivity, network states can hold prior values without decay or growth, allowing integration of inputs over time. Thus, this mechanism is called a robust integrator. Energy wells can be spaced arbitrarily close together while maintaining their depth, so that the robust integrator can represent a practically continuous range of values. However, the energy wells imply a minimum input strength to transition between adjacent states, with inputs below this limit effectively ignored. 2 (A) (B) Figure 1: Schematic of neural integrator models. (A) Visualizing integration via an energy surface (Pouget and Latham, 2002). A robust integrator can "fixate" at a range of discrete values, indicated by a sequence of potential wells, despite mistuning of circuit feedback. Without these wells (the non-robust case), activity in a mistuned integrator would either exponentially grow or decay, as in the top panels. Perturbing the robust integrator from one well to the next, however, requires sufficiently strong momentary input. (B) As a consequence, low-amplitude segments in the input signal ∆I(t), below a robustness limit R, are not accumulated by a robust integrator: only the high- amplitude segments are. The piecewise-defined differential Equation (5) captures this robustness behavior, resulting in the accumulated activity shown, and may be related to, e.g., a detailed bistable-subpopulation model. A decision is expressed when the accumulated value E(t) crosses the decision threshold θ. The two models just introduced present a tradeoff between robustness to parameter mistun- ing and sensitivity to inputs. Here, we ask how this tradeoff impacts behavioral performance in perceptual decision making. Focusing on the moving dots task (Shadlen and Newsome, 1996; Roitman and Shadlen, 2002), enables us to constrain model parameters to known physiology and behavior. Our aim is to establish whether or not the robust integrator model is consistent with known data, and to assess the performance benefits, if any, that it affords when network parameters cannot be fine-tuned. 2 Materials and methods 2.1 Model and task overview To explore the consequences of the robust integrator mechanism for decision performance, we begin by constructing a two-alternative decision making model similar to that proposed by Mazurek et al. (2003). For concreteness, we concentrate on the forced choice motion discrimination task (Roit- man and Shadlen, 2002; Mazurek et al., 2003; Gold and Shadlen, 2007; Churchland et al., 2008; Shadlen and Newsome, 1996; Shadlen and Newsome, 2001). Here, subjects are presented with a field of random dots, of which a subset move coherently in one direction; the remainder are relo- cated randomly in each frame. The task is to correctly choose the direction of coherent motion from two alternatives (i.e., left vs. right). As in Mazurek et al. (2003) (see also Smith (2010)), we first simulate a population of neurons that represent the sensory input to be integrated over time. This population is a rough model of cells in extrastriate cortex (Area MT) which encode momentary information about motion direc- tion (Britten et al., 1993; Britten et al., 1992; Salzman et al., 1992). We pool spikes from model 3 Not Robust: Perfect Over-tuned Under-tuned: Robust: Perfect Over-tuned Under-tuned: Figure 2: Overview of model setup. Simulations of sensory neurons and neural recordings are used to define the left and right inputs ∆Il(t), ∆Ir(t) to neural integrators (see text). These inputs are modeled by gaussian (OU) processes, which capture noise in the encoding of the motion strength by each pool of spiking neurons. See Equations (1)-(3) for definition of input signals. Similar to Mazurek et al. (2003), the activity levels of the left and right integrators El(t) and Er(t) encode In the reaction time task, El(t) and Er(t) race to accumulated evidence for each alternative. thresholds in order to determine choice on each trial. In the controlled duration task the choice is made in favor of the integrator with higher activity at the end of the stimulus presentation. MT cells that are selective for each of the two possible directions into separate streams, labeled according to their preferred "left" and "right" motion selectivity: see Figure 2. Two corresponding integrators then accumulate the difference between these streams, left-less- right or vice-versa. Each integrator therefore accumulates the evidence for one alternative over the other. Depending on the task paradigm, different criteria may be used to terminate accumulation and give a decision. In the reaction time task, accumulation continues until activity crosses a decision threshold: if the leftward evidence integrator reaches threshold first, a decision that overall motion favored the leftward alternative is registered. Accuracy is defined as the fraction of trials that reach a correct decision. Speed is measured by the time taken to cross threshold starting from stimulus onset. Reaction Time (RT) is then defined as the time until threshold (decision time) plus 350 ms of non-decision time, accounting for other delays that add to the time taken to select an alternative (e.g. visual latencies, or motor preparation time, cf. (Mazurek et al., 2003; Luce, 1986)). The exact value of this parameter was not critical to our results. Task difficulty is determined by the fraction of coherently moving dots C (Britten et al., 1992; Mazurek et al., 2003; Roitman and Shadlen, 2002). Accuracy and RT across multiple levels of task difficulty define the accuracy and chronometric functions in the reaction time task, and together can be used to assess model performance. When necessary, these two numbers can be collapsed into a single metric, such as the reward per unit time or reward rate. In a second task paradigm, the controlled duration task, motion viewing duration is set in advance by the experimenter. A choice is made in favor of the integrator with greater activity at the end of the stimulus duration. Here, the only measure of task performance is the accuracy function. 4 Sensory Neurons: Input Signals: Recurrent Integration: Threshold: Left: Right: Figure 3: Construction of gaussian (OU) processes to represent fluctuating, trial-by-trial firing rate of a pool of weakly correlated MT neurons (Bair et al., 2001; Zohary et al., 1994). As in Mazurek and Shadlen (2002), these motion sensitive neurons provide direct input to our model integrator circuits. Simulated spike trains from weakly-correlated, direction selective pools of neurons are shown as a rastergram. All spikes prior to time t -- a sum over the jth spike from the ith neuron, for all i and j -- are convolved with an exponential filter, and then summed to create a continuous stochastic output (right); here, H(t) is the Heaviside function. We approximated this output by a simpler gaussian (OU) process in order to simplify numerical and analytical computations that follow. 2.2 Sensory input We now describe in detail the signals that are accumulated by the integrators corresponding to the "left" and "right" alternatives. First, we model the pools of leftward or rightward direction- selective sensory (MT) neurons as N = 100 weakly correlated spiking cells (Pearson's correlation ρ = .11 (Zohary et al., 1994; Bair et al., 2001)); see Figure 3. Specifically, as in Mazurek and Shadlen (2002), each neuron is modeled via an unbiased random walk to a spiking threshold; the random walks of neurons in the same pool are correlated. Increasing the variance of each step in the random walk increases the firing rate of each model neuron; it was therefore chosen at each coherence value to reproduce the linear relationship between coherence C and mean firing rate µl,r of the left and right selective neurons observed in MT recordings: µl,r(C) = r0 + bl,rC . (1) Here the parameters r0, bl, and br are derived from firing rates observed across a range of coher- ences (Britten et al., 1993). If evidence favors the left alternative, bl = .4 and br = −.2; if the right alternative is favored, these values are exchanged. Next, the output of each spiking pool was aggregated. Each spike emitted from a neuron in the pool was convolved with an exponential filter with time constant 20 ms. This is intended as an approximate model of the smoothing effect of synaptic transmission. These smoothed responses were then summed to form a single stochastic process for each pool (see Figure 3, right). We then approximated the smoothed output of each spiking pool by a simpler stochastic process that captures the mean, variance, and temporal correlation of this output as a function of dot coherence. We used gaussian processes Il(t) and Ir(t) for the rightward- and leftward-selective pools (See Figure 3). Specifically, we chose Ornstein-Uhlenbeck (OU) processes, which are continuous 5 Cell # Sensory Neurons gaussian process generated by the stochastic differential equations dIl,r = µl,r(C) − Il,r τ dt +r 2νl,r(C) τ dWt (2) with mean µl,r(C) as dictated by Equation 1. The variance νl,r(C) and timescale τ were chosen to match the steady-state variance and autocorrelation function of the smoothed spiking process. As we will see, this timescale plays an important role in determining the decision making performance of robust integrators. Our construction so far accounts for variability in output from left vs. right direction selective neurons. We now incorporate an additional noise source into the output of each pool. These noise terms (ηl(t) and ηr(t), respectively) could approximate, for example, neurons added to each pool that are nonselective to direction. Each noise source is modeled as an independent OU process with mean 0, timescale 20 ms as above, and a strength (variance) νγ/2. This noise strength is a free parameter that we vary to match behavioral data (see "A robust integrator circuit" and Figure 14). We note that previous studies also found that performance based on the direction-sensitive cells alone can be more accurate than behavior, and therefore incorporated variability in addition to the output of "left" and "right" direction selective MT cells (Shadlen et al., 1996; Mazurek et al., 2003; Cohen and Newsome, 2009). Finally, the signals that are accumulated by the left and right neural integrators are constructed by differencing the outputs of the two neural pools: ∆Il(t) = [Il(t) + ηl(t)] − [Ir(t) + ηr(t)] ∆Ir(t) = −∆Il(t) . (3) 2.3 Neural integrator circuit and feedback mistuning A central focus of our paper is variability in the relative tuning of recurrent feedback vs. decay in an integrator circuit. Below, we will introduce the mistuning parameter β, which determines the extent to which feedback and decay fail to perfectly balance. We first define the dynamics of the integrator circuit on which our studies are based. This is described by the firing rates El,r(t) of integrators that receive outputs from left-selective or right-selective pools ∆Il,r(t) respectively. The firing rates El,r(t) increase as evidence for the corresponding task alternative is accumulated over time: dEl,r dt τE = −El,r + (1 + β)El,r + κ∆Il,r(t). (4) The three terms in this equation account for leak, feedback excitation, and the sensory input (scaled by a weight κ), respectively. When the mistuning parameter β = 0, leak and self-excitation exactly cancel; we describe such an integrator as perfectly tuned, while an integrator with β 6= 0 is said to be mistuned. Imprecise feedback tuning is modeled by randomly setting β to different values from trial to trial (but constant during a given trial), with a mean value ¯β and a precision given by a standard deviation σβ. We assume that ¯β = 0 for most of the study. Thus the spread of β, which we take to be gaussian, represents the intrinsic variability in the balance between circuit-level feedback and decay. Perfect tuning corresponds to σβ = ¯β = 0, while σβ 6= 0 or ¯β 6= 0 corresponds to a mistuned integrator. Finally, we set initial activity in the integrators to zero (El,r(0) = 0), and impose reflecting boundaries at Er = 0, El = 0 (as in, e.g., Smith and Ratcliff (2004)) so that firing rates never become negative. 6 2.4 A robust integrator circuit A robust integrator can be constructed from a series of bistable subpopulations, which sequentially activate in order to represent the accumulated evidence (Koulakov et al., 2002; Nikitchenko and Koulakov, 2008). The many equations that describe the evolution of these systems can be closely approximated with reduced models, as demonstrated in Goldman et al. (2003). We derived a single piecewise-defined differential equation model that approximates the dynamics of a robust integrator constructed from bistable pools. All subsequent results are based on this simplified model, which captures the essence of the robust integration computation: τE dEl,r dt =(cid:26) 0 βEl,r + κ∆Il,r : βEl,r + κ∆Il,r ≤ R : otherwise (5) The first line represents the series of potential wells discussed in the Introduction (see Figure 4): if the sum of the mistuned integrator feedback and the input falls below the robustness limit R, the activity of the integrator remains fixed. If this summed input exceeds R, the activity evolves as for the non-robust integrator in Equation (4). To interpret the robustness limit R, it is convenient to normalize by the standard deviation of the input signal: R = R ST D [∆Il,r(t)] . (6) In this way, R can be interpreted in units of standard deviations of input OU process that are "ignored" by the integrator. We note that Equation (5) is similar to the effective equation derived for a different implementation of a robust integrator (Goldman et al., 2003). To summarize, Equation (5) defines a parameterized family of neural integrators, distinguished by the robustness limit R. As R → 0, the model reduces to Equation (4). When additionally β = 0, the (perfectly tuned) integrator computes an exact integral of its input: Equation (5) then yields El,r(t) ∝R t 0 ∆Il,r(t0)dt0. 2.5 Computational methods Monte Carlo simulations of Equations (1)-(5) were performed with Euler-Maruyama method (Higham, 2001), with dt = 0.1 ms. For a fixed choice of input statistics and threshold θ, a minimum of 10, 000 trials were simulated to estimate accuracy and RT values. During simulations of the reaction time task, in order to prevent excessively long trials (particularly at low coherence values) a maximum simulation time was set at 10,000 ms. At this time, if neither integrator had reached threshold, the indeterminate result was broken by a numerical "coin flip", (this rarely occurred, as indicated by the RT histograms in Figures 15, 16). In simulations where σβ > 0, results were generated across a range of β values and then marginalized by weighting according to a normal distribution. The range of values was chosen with no less than 19 linearly spaced points, across a range of ± 3 standard deviations around the mean ¯ β. Reward rate values presented in "Reward rate and the robustness-sensitivity tradeoff" are pre- sented as maximized by varying the free parameter θ; values were computed by simulating across a range of θ values. The range and spacing of these values were chosen dependent on the values of R and β for the simulation; the range was adjusted to capture the relative maximum of reward rate as a function of θ, while the spacing was adjusted to find the optimal θ value with a resolution of 7 Figure 4: Parameter space view of four integrator models, with different values of the robustness limit R and feedback mistuning σβ. The impact of transitioning from one model to another by changing parameters is either to enhance or diminish performance, or to have a neutral effect (see text). ± 0.1. Values of θ and νγ in the table included in Figure 14 were chosen to best match accuracy and chronometric functions to behavioral data reported in Roitman and Shadlen (2002). This was accomplished by minimizing the sum-squared error in data vs. model accuracy and chronometric curves across a discrete grid of θ and νγ values, with a resolution of 0.1. When data between simulated values were needed, linear interpolation was used to approximate the corresponding accuracy and RT values. Autocovariance functions of integrator input, presented later, were computed by simulating an Ornstein-Uhlenbeck process using the exact numerical technique in Gillespie (1996) with dt = 0.1 ms, to obtain a total of 227 sample values. Sample values of the process less than the specified robustness limit R were set to 0, and the autocovariance function was computed using standard Fourier transform techniques. Simulations were performed on NSF Teragrid clusters. 3 Results 3.1 How do robustness and mistuning affect decision speed and accuracy? In the Methods, we define a general neural integrator model (Equation (5)) that accumulates signals representing the output of motion sensitive neurons (Equation 3). The integrator model includes two key parameters. The first is β, which represents the mistuning of feedback from a value that perfectly balances decay; the extent of this mistuning is measured by σβ, the standard deviation of β from the ideal value β = 0. The second is the robustness limit R. We emphasize twin effects of R: as R increases, the integrator becomes able to produce a range of graded persistent activity for ever-increasing levels of mistuning (see Figure 4 (A), where R corresponds to the depth of energy wells). This prevents runaway increase or decay of activity when integrators are mistuned; intuitively, this might lead to better performance on sensory accumulation tasks. At the same time, as R increases integrator activity remains fixed even for increasingly strong positive or negative momentary input ∆Il,r (see Figure 4 (B), where R specifies a limit within which inputs are "ignored"). Such sensitivity loss should lead to worse performance. This implies a fundamental tradeoff between competing effects: (1) one would prefer to not ignore relevant input stimulus, 8 "Mistuned" Model "Recovery" Model "Baseline" Model "Robust" Model Figure 5: Mistuned feedback diminishes decision performance. (Inset) In both figures we consider a move in parameter space from the "baseline" model to the "mistuned" model by changing σβ = 0 → 0.1 (A) In the controlled duration task, accuracy is lower for the "mistuned" model (dashed line) than for the "baseline" model (solid line) at every trial duration T , indicating a loss of performance when σβ increases. (B) In the reaction time task, we plot the curve of all (RT, accuracy) pairs attained by varying the decision threshold θ (see text). Once again, accuracy is diminished by mistuning. favoring small R, and (2) one would prefer an integrator robust to mistuning, favoring large R. Figure 4 gives a schematic of how the two model parameters, σβ and R, define a plane of possible integrator models. Here, we explore decision performance in four different cases arranged in this plane. By contrasting integrators with different values of the robustness limit R, we can assess how the fundamental tradeoff plays out, to either improve or degrade decision making performance. In order to assess this performance, we consider relationships between decision speed and accu- racy in both controlled duration and reaction time tasks. In the controlled duration task, we simply vary the stimulus presentation duration, and plot accuracy vs. experimenter-controlled stimulus duration. In the reaction time task, we vary the decision threshold θ -- treated as a free parameter -- over a range of values, thus tracing out the speed accuracy curve for all possible pairs of speed and accuracy values. Here, speed is measured by reaction time (RT), the latency between the onset of stimulus and crossing of the decision threshold. For both cases, we use a single representative dot coherence (C=12.8 in Equation 1); results are qualitatively similar for other coherence values (data not shown); slightly (approx. 25%) lower robustness limits are required at the lowest dot coherence of C = 3.2. We first study a case we call the "baseline" model, for which there is no mistuning or robustness: σβ = R = 0. Speed accuracy plots for this model are shown as a solid line in Figs. 5(A) and (B), for the controlled duration and reaction time tasks respectively. We compare the "baseline" model with the "mistuned" model, for which the feedback parameter has a standard deviation of σβ = 0.1 (10% of the mean feedback) and robustness R = 0 remains unchanged. In the controlled duration task (Figure 5(A)) we observe that mistuning diminishes accuracy by as much as 10%, and this effect is sustained even for arbitrarily long viewing windows (cf. (Usher and McClelland, 2001; 9 Controlled Duration (sec) Reaction Time (ms) Accuracy Accuracy Figure 6: Increasing the robustness limit R helps recover performance lost due to feedback mis- tuning. (Inset) We illustrate this by moving in parameter space from the "mistuned" model to the "recovery" model, by changing R = 0 → .85. The impact on decision performance is shown for both the controlled duration (A) and reaction time (B) tasks. For each task we plot the rela- tionship between speed and accuracy as above: solid lines indicate the "baseline" model, dotted the "mistuned" model, and now dash-dotted the "recovery" model. We find that R > 0 yields a modest performance gain for the "recovery" model in comparison with the "mistuned" model. Bogacz et al., 2006)). The reaction time task (Panel B) produces a similar effect: for a fixed RT, the corresponding accuracy is decreased. Next we increase the robustness limit to R = 0.85 -- so that almost ± a standard deviation of the input stream is "ignored" by the integrators -- while maintaining feedback mistuning. We call this case the "recovery" model because robustness compensates in part for the performance loss due to feedback mistuning: the speed accuracy plots in Figure 6 for the recovery case lie above those for the "mistuned" model. For example, at the longer controlled task durations (Panel A) and reaction times (Panel B) plotted, 20% of the accuracy lost due to integrator mistuning is recovered via the robustness limit R = 0.85. Finally, we study the remaining possibility, when the robustness parameter R is increased from zero in a perfectly tuned integrator (σβ = 0); this is the "robust" case in Figure 4. We expected performance to be substantially diminished as a consequence of lost sensitivity to inputs. However, Figure 7 demonstrates that this is not the case: speed accuracy curves for R = 0.85 almost coincide with those for the "baseline" case of R = 0. We emphasize again that because R measures ignored input in units of the standard deviation, the integrator circuit is actually not integrating the weakest 60% of the input stimulus. Given this large amount of ignored stimulus, the fact that the robust integrator produces nearly the same accuracy and speed as the "baseline" case is surprising. This implies that the "robust" model can protect against feedback mistuning, without substantially sacrificing performance when feedback is perfectly tuned. To summarize, the ratchet-like mechanism of the robust integrator appears well-suited to the decision tasks at hand. This mechanism counteracts some of the performance lost when feedback fails to be perfectly fine-tuned. Moreover, even when this fine-tuning is achieved, a robust integrator 10 Controlled Duration (sec) Reaction Time (ms) Accuracy Accuracy Figure 7: Increasing R alone does not compromise performance. (Inset) We illustrate this by moving in parameter space from the "baseline" to the "robust" model. For both (A) controlled duration and (B) reaction time tasks, we plot the relationship between speed and accuracy. Solid lines give results for the "baseline" model, and dash-dotted for the robust model. The curves are very similar in the "baseline" and robust cases, indicating little change in decision performance due to the robustness limit R = 0.85. still performs as well as the "baseline" case that is perfectly sensitive to the input signal. In the next section, we begin to explain this observation by constructing several simplified models and employing results from statistical decision making theory. 3.2 Analysis: Robust integrators and decision performance 3.2.1 Controlled duration task: Discrete time analysis We can begin to understand the effect of the robustness limit on decision performance by formu- lating a simplified version of the evidence accumulation process. We focus first on the controlled duration task, where the analysis is somewhat simpler. Our first simplification is to consider a single accumulator E which receives evidence for or against a task alternative in discrete time. The value of E on the ith time step, Ei, is allowed to be either positive or negative, corresponding to accumulated evidence favoring the leftward or rightward alternatives, respectively. On each time step, Ei increments by an independent, random value Zi with a probability density function (PDF) fZ(Z). We first describe an analog of the "baseline" model above; i.e., in the absence of robustness ( R = 0). Here, we take the increments Zi to be independent, identical, and normally distributed, with a mean µ > 0 (i.e. biased toward the leftward alternative; we call this the preferred alternative) and standard deviation σ: that is, Zi ∼ N(cid:0)µ, σ2(cid:1). After the nth step, we have En = Zi . n X i=1 11 Controlled Duration (sec) Reaction Time (ms) Accuracy Accuracy (A) (B) Figure 8: The effect of R on a discrete time increment distribution, and the second real root of the moment generating function of this distribution. (A) The PDF of the random variable Z R, with probability mass for values between the robustness limit R re-allocated as a delta function centered at zero ( R = 1). (B) The second real root h0 of MZ R (s) remains unchanged as R increases from 0 → 2. (Lines are uniformly distributed in this range.) This implies that in the reaction time task, no changes in the accuracy and chronometric functions will be observed until the deviation in E[ZR] becomes large (Discussed in "Reaction time task: Continuous analysis"). In the controlled duration task, a decision is rendered after a fixed number of time steps N , (i.e. n = N ) and a correct decision (i.e., in favor of the preferred alternative) occurs when EN > 0. By construction, En ∼ N(cid:0)nµ, nσ2(cid:1), which implies that accuracy (Acc) can be computed as a function of the signal-to-noise ratio (SNR) s = µ σ of a sample: Acc =Z ∞ 0 1 √2πN σ2 e− (x−N µ)2 2N σ2 dx = 1 + Erf(cid:18)q N 2 s(cid:19) 2 . (7) Next, we change the distribution of the accumulated increments Zi to construct a discrete time analog of the robust integrator. Specifically, increasing the robustness parameter to R > 0 affects increments Zi by redefining the PDF fZ(Z) so that weak samples do not add to the total accumulated "evidence", precisely as in Equation (5). (Models where such a central "region of uncertainty" of the sampling distribution is ignored have previously been studied in a race-to- bound model (Smith and Vickers, 1989); see Discussion). This requires reallocating probability mass below the robustness limit to zero. We plot the resulting PDF in Figure 8(A), where the reallocated mass gives a weighted delta function at zero. Specifically: fZR (Z) = δ (0)Z R −R fZ(cid:0)Z0(cid:1) dZ0 +(cid:26) 0 fZ(Z) Z < R : : otherwise (8) The central limit theorem then allows us to approximate the new cumulative sum EN R as a normal distribution (for sufficiently large N ), with µ and σ in Equation (7) replaced by the mean and standard deviation of the PDF defined by Equation (8). As before, we normalize R by the standard deviation of the increment, R = R σ , and then express the fraction correct Acc R as a function of R and s. One can think of R as perturbing the original Acc function (Equation (7)), and although this perturbation has a complicated form, we can understand its behavior by 12 Figure 9: Comparison of the decision accuracy predicted by discrete and continuous time approxi- mate models for the controlled duration task (as above, coherence C = 12.8). (A) Both discrete and continuous time models predict that accuracy (Acc) will stay roughly constant as R increases up to ≈ 0.5, followed by a gradual decrease. However, the continuous time approximation provides a closer match to results for the full model pictured in Figure 7(A) (see text). Additionally, compar- ing the dashed and dot-dashed lines shows that the approximation in Equation 9 provides a good description of the discrete time model for R . 1. (B) The (numerically computed) autocovariance functions of the input signal Z R(t) at various levels of R that were used to construct the continuous time curve in (A) (Equation (9)). (Inset) Two of these same autocovariance functions (for R = 0 and R = 1.5) are plotted normalized to their peak value. This shows that autocorrelation falls off faster as R increases. observing that its Taylor expansion has only one nonzero term up to fifth order in R: Acc R(N ) ≈ Acc(N )(1 − P (N )) (9) P (N ) = √N s(cid:0)1 + 2s2(cid:1) e− (1+N )s2 6π 2 R3 + O( R5) Thus, for small values of R (giving very small R3), there will be little impact on accuracy. Equation (9) can therefore partially explain the key observation in Fig 7(A) that R can be increased to 0.85 while incurring very little performance loss. For concreteness, we focus on decisions at T = 500 ms. To make a rough comparison, we first assume that a new sample of evidence arrives in the discrete time model every 10 ms. We then set the SNR s so that Acc = 0.92 for the discrete model when N = 50, agreeing with the accuracy obtained from the continuous model at 500 ms (Fig 7(A)) when R = 0. We then increase the robustness limit R. Figure 9(A) shows that accuracy for both the discrete time model itself (dashed line), and its approximation up to O( R5) (dot-dashed), barely decrease at all while R is less than ≈ 0.5, and then begin to fall off. This is consistent with the results for the full model in Fig 7(A). However, the discrete time model does predict a small decrease in accuracy at R = 0.85 that is not seen in the full model. In the next section, we explain how this discrepancy can be resolved. 3.2.2 Controlled duration task: Continuous time analysis We next extend the analysis of the controlled duration task in the previous section to signal inte- gration in continuous time. In brief, we follow a method developed in Gillespie (1996) to describe the evolution of the mean and variance of a continuous input signal that has been integrated over 13 Accuracy time. This is challenging and interesting because, as for the signals used in modeling the random dots task above (see Methods, Sensory Input), this input signal contains temporal correlations. As in the previous section, we describe the distribution of the integrated signal at the final time T , which determines accuracy in the controlled duration task. We first replace the discrete input samples Zi from the previous section with a continuous signal Z(t), which we take to be a (OU) gaussian process with a correlation timescale derived from our model sensory neurons (see Methods). We define the integrated process dE dt = Z(t) → E(t) =Z t 0 Z(t0)dt0 with initial condition E(0) = 0. Assuming that Z(t) satisfies certain technical conditions that are easily verified for the OU pro- cess (wide-sense stationarity, α-stability, and continuity of sample paths (Gardiner, 2002; Billings- ley, 1986; Gillespie, 1996)), we can construct differential equations for the first and second moments hE(t)i and(cid:10)E2(t)(cid:11) evolving in time. We start by taking averages on both sides of our definition of E(t), and, noting that E(0) = 0, compute the time-varying mean: dhE(t)i dt = hZ(t)i =⇒ hE(t)i = thZ(t)i . (10) Similarly, we can derive a differential equation for the second moment of E(t): d(cid:10)E2(t)(cid:11) dt = 2hZ(t)E(t)i . The righthand side of this equation can be related to the area under the autocovariance function A (τ ) ≡ hZ(t)Z(t + τ )i − hZ(t)i2 of the process Z(t): hZ(t)E(t)i =(cid:28)Z(t)Z t 0 Z(s)ds(cid:29) = Z t = Z t = Z t 0 0 hZ(t)Z(s)i ds 0 hZ(t)Z(t − τ )i dτ A (τ ) + hZ(t)i2 dτ We now have an expression for how the second moment evolves in time. We can simplify the result via integration by parts: (cid:10)E2(t)(cid:11) = 2Z t 0 Z s 0 A (τ ) + hZ(t)i2 dτ ds = 2Z t 0 (t − τ )A (τ ) dτ + t2 hZ(t)i2 =⇒ Var[E(t)] = 2Z t 0 (t − τ )A (τ ) dτ. (11) Because E(t) is an accumulation of gaussian random samples Z(t), it will also be normally dis- tributed, and hence fully described by the mean (Equation (10)) and variance (Equation (11)) (Billings- ley, 1986). To model a non-robust integrator, as discussed above we take Z(t) to be a OU process with steady-state mean and variance µ and σ2, and time constant τ . For the robust case, we can follow Equation 5 and parameterize a family of processes Z R(t) with momentary values below 14 the robustness limit R set to zero. (Here, we again normalize the robustness limit by the standard deviation of the OU process.) We numerically compute the autocovariance functions A R (τ ) of these processes, and use the result to compute the required mean and variance, and hence time-dependent signal-to-noise ratio SNR(t), for the integrated process E(t). This yields SN R R(t) = . (12) hE(t)i pV ar[E(t)] = q2R t tE[Z R(t)] 0 (t − τ )A R (τ ) dτ Under the assumption that E(T ) is approximately gaussian for sufficiently long T (which can be verified numerically), we use this SNR to compute decision accuracy at T : 1 + Erf(cid:16)q 1 2 SN R R(T )(cid:17) 2 Acc R(T ) ≈ . (13) This function is plotted for T = 500 ms as the solid line in Figure 9(A). The plot shows that accuracy remains relatively constant until the robustness limit R exceeds ≈ 0.85. Interestingly, this is a longer range of R values than for the discrete time case (compare dotted line in Figure 9(A)), and is closer to the results for the full model pictured in Figure 7(A). Why does the robustness limit appear to have a milder effect on degrading decision accuracy for our continuous vs. discrete time input signals? We can get some insight into the answer by examining the autocovariance functions A R (τ ), which we present in Figure 9(B). When normalized by their peak value, the autocovariance for R > 0.5 falls off more quickly vs. the time lag τ (see inset in Figure 9(B)), indicating that subsequent samples become less correlated in time. Thus, there are effectively more "independent" samples that are drawn over a given time range T , improving the fidelity of the signal. This effect is not present in our discrete time model. Summary: Our analysis of decision performance for the controlled duration task shows that two factors contribute to the preservation of decision performance for robust integrators. The first is that, for robustness limits up to R ≈ 0.5, the momentary SNR of the inputs is barely changed by setting values below robustness limit to zero. The second is that, as R increases, the signal Z R(t) being integrated becomes less correlated in time. This means that (roughly) more independent samples will arrive over a given time period. 3.2.3 Reaction time task: Discrete analysis We begin our analysis of the reaction time task by introducing a discrete time, discrete space random walk model. In this model, schematized in Figure 10(A) with five intermediate states, a particle representing the accumulated value E starts at a state balanced between two absorbing "sink" states. At every time step, the particle moves towards the "correct" (i.e. preferred) sink with probability p(1−R), and towards the "incorrect" (null) sink with probability (1−p)(1−R) (we consider p > 0.5, biasing the random walk toward the "correct" sink). There is also the possibility that the particle might remain in the current state, with probability R. We now draw an analogy between the states in this random walk model and the ratcheting dynamics among energy wells in a robust integrator (see Figure 4 and Introduction). Here, the po- sition of the particle represents the accumulated evidence for the left vs. right alternatives, and the absorbing states represent crossing of the corresponding decision thresholds. When the robustness limit R is increased, the wells -- each of which could represent a bistable neural subpopulation (see Methods) -- act to hold the particle in a given state, with a probability set by R. 15 (A) (B) Figure 10: Analysis of the effect of the robustness limit in the reaction time task. (A) Biased random walk between two absorbing boundaries. The correct and incorrect states act as "sinks" of the discrete time discrete space Markov chain. The intermediate states are analogous to the potential wells in the robust integrator model. Here, a particle will remain in the current state at the next time step with probability R. The final probability of ending up in either sink is independent of R. (B) Speed accuracy tradeoff curves from Figure 7 are replotted and labeled (c), while a new line labeled (d) shows the performance predicted by the discrete time, continuous space model described in "Reaction time task: Continuous analysis". Here the signal-to-noise ratio of the discrete increments were chosen so that the line generated in the R = 0 case would overlay the line from the R = 0 continuous model, and is therefore not plotted. We see that the discrete time model over-predicts the impact of the robustness limit R, just as in the controlled duration case (Figure 9). 16 Reaction Time (ms) Accuracy As R is increased in the random walk model, the probability of transitioning out of a given state similarly decreases. Standard results on Markov chains (see, for example, Kemeny and Snell (1960)) provide formulas for the probability that the particle will end in one vs. the other sink, as well as the expected number of time steps until this occurs, based on the transition matrix associated with the random walk. The probability of ending in the "correct" sink corresponds to decision accuracy, and is found at the middle entry in the solution vector x of the matrix equation (I − Q) x = (1 − R)pe1 . (14) Here Q is a tridiagonal matrix with R on the main diagonal, p(1 − R) on the lower diagonal, and (1 − p)(1 − R) on the upper diagonal; e1 is the canonical basis vector with e(1) 1 = 1, and all other entries equal to 0. After some factoring, we find a common factor of (1 − R) on both sides of the equation; thus the solution to x is independent of R. This implies that the probability of ending up in the correct state is unchanged by increasing R from the non-robust case (R = 0). Intuitively this makes sense: if one conditions on the fact that one will leave the current state on the next time step, the probability of moving toward the correct and incorrect states are independent of R. The same is not true for the expected number of steps necessary to reach a sink (by analogy, the reaction time). This is because the matrix system that yields reaction times is: (I − Q) t = 1. (15) Here the right-hand side of this equation is the vector of all ones, and therefore no equivalent cancellation can occur. However, we do notice that the reaction time with R 6= 0 is just a rescaling of the original reaction time with R = 0. Specifically, if tR is the expected number of steps required to reach an absorbing state, then tR = t0 . (16) 1 1 − R (17) (18) h0(cid:21) θ 2 Acc = RT = 1 1 + eθh0 θ E[Z] tanh(cid:20)− 17 Summary: We have used a simplified random walk model to gain intuition about the effect of the robustness limit in the reaction time task, and to show that adding a robustness limit only affects decision latency, but not accuracy. In the next section, we will derive a similar result for continuous sample distributions. 3.2.4 Reaction time task: Continuous analysis We return to the continuous sampling distribution introduced in "Controlled duration task: Dis- crete time analysis", but now in the context of threshold crossing in the reaction time task. The accumulation of these increments toward decision thresholds can be understood as the sequen- tial probability ratio test, where the log-odds for each alternative are summed until a predefined threshold is reached (Wald, 1945; Gold and Shadlen, 2002; Luce, 1963; Laming, 1968). Wald (1944) provides an elegant method of computing decision accuracy and speed (RT). The key quantity is given by the moment generating function (MGF, denoted MZ(s) and defined in Equation 19) for the samples Z (see Luce (1986) and Doya (2007), Chapter 10). Under the assumption that thresholds are crossed with minimal overshoot, we have the following expressions: Thus, the only effect of the robustness limit R is to delay arrival at the sinks. where h0 is one of the two real roots of the equation MZ(s) = 1 (the other root is precisely 1) and θ is the decision threshold. We first consider the case of a non-robust integrator, for which the samples Z are again normally distributed. In this case, we must solve the following equation to find s = h0: MZ(s) = EZ [es∗z] =Z ∞ It follows that s = 1 and s = h0 = −2 µ σ2 provide two real solutions of this equation. (Wald's Lemma ensures that there are exactly two such real roots, for any sampling distribution meeting easily satisfied technical criteria.) fZ (z) es∗zdz = e s2σ2 2 +sµ = 1. (19) −∞ When the robustness limit R > 0, we can again compute the two real roots of the associated MGF. Here, we use the increment distribution fZR(Z) given by Equation (8), for which all proba- bility mass within R of 0 is reassigned precisely to 0. Surprisingly, upon plugging this distribution into the expression MZ(s) = 1, we find that s = 1, h0 continue to provide the two real solutions to this equation regardless of R, as depicted in Figure 8(B). This observation implies that (1) accuracies (Equation (17)) are unchanged as R is increased, and (2) reaction times (Equation (18)) only change when E[ZR] changes. In other words, the inte- grator can ignore inputs below an arbitrary robustness limit at no cost to accuracy, and a penalty in terms of reaction time will only be observed when E[ZR] changes appreciably. Generalizing our result, we note that a sufficient condition for h0 to be unchanged as R changes is that the original sampling distribution fZ(z) obeys fZ(z) = fZ(−z)eh0z; it is straightforward to verify that the Gaussian satisfies this property. We next determine the magnitude of R necessary to change E[ZR]. When we substitute R = R σ and compute the perturbation to E[ZR], we again find only one term up to fifth order in R: (20) (21) E[ZR] =Z ∞ P =r 2 −∞ 9π z fZR (z) dz = E[Z] (1 − P ) 2 ( µ σ )2 e− 1 R3 + O( R5) This outcome is similar to the controlled duration case: small values of R will have little effect on E[ZR], and therefore little effect on increasing decision speed (via Equation (18)). Moreover, as we have already shown, accuracy is unaffected by robustness limits R of any value. As a consequence we expect speed accuracy curves to change only modestly for small values of R. We illustrate this via a speed accuracy plot in Figure 10(B). Here, the present discrete time, continuous space model produces the chain-dotted curve (marked (d)), showing a moderate decrease in performance at R = 0.85. This decrease is purely due to the increase in RT just discussed. However, the model at hand does not reproduce the speed accuracy curve for the continuous time model shown in Figure 10(B). Indeed, the continuous time model produces better performance (higher accuracy at a given speed). This suggests an additional effect in the continuous time case: once again, the fact that R reduces autocorrelation of the integrated signal increases the fidelity of the input, improving performance (see inset in Figure 9(B)). Unlike the simpler controlled duration task, attempting a mathematical analysis of this effect is beyond the scope of this paper. Summary of analysis: We pause to summarize our analysis of how the robustness limit R impacts decision performance. For both the controlled duration and reaction time tasks, we first studied 18 Figure 11: Using the Reward Rate metric to quantify recovery of decision performance as the robustness limit is increased, C = 12.8. (A) Speed accuracy curves plotted for multiple values of R; as in previous figures, the greater accuracies found at fixed reaction times indicate that performance improves as R increases. The heavy line indicates the "baseline" case of a perfectly tuned, non-robust integrator (repeated from Figure 5(B)). RR isoclines are plotted in background (dotted lines; see text), and points along speed accuracy curves that maximize RR are shown as circles. These maximal values of reward rate are plotted in (B), demonstrating the non-monotonic relationship between R and the best achievable RR. the effect of this limit on the evidence carried by momentary values of sensory inputs. In each task, this effect was more favorable than might have been expected: in the controlled duration case, the signal to noise ratio of momentary inputs was preserved for a fairly broad range of R, while in the reaction time task, R was shown to affect accuracy but not speed at fixed decision threshold. Moreover, the robustness mechanism serves to decorrelate input signals in time. This contributes further to decision performance being preserved as the robustness limit increases. 3.3 Reward rate and the robustness-sensitivity tradeoff Until now, we have examined performance in the reaction time task by plotting the full range of attainable speed and accuracy values. The advantage of this approach is that it demonstrates decision performance in a general way. An alternative, more compact approach, is to assume a specific method of combining speed and accuracy into a single performance metric. This approach is useful in quantifying decision performance, and rapidly comparing a wide range of models. Specifically, we use the reward rate (RR) (Gold and Shadlen, 2002; Bogacz et al., 2006): RR = F C RT + Tdel . (22) Reward rate can be thought of as the number of correct responses made per unit time, with a delay Tdel imposed between responses to penalize rapid guessing. Implicitly, this assumes a motivation on the part of the subject which may not be true; in general, subjects rarely achieve 19 0.0 .6 .85 1.15 Reaction Time (ms) Accuracy (A) (B) Figure 12: A nonzero robustness limit improves performance across a range of mistuning baises ¯ β. In both the reaction time (A) and controlled duration (B) tasks, robustness helps improve performance when β ∼(cid:0) ¯β, .12(cid:1), for all values of ¯β shown. Here, as in previous panels, the coherence of the sensory input is C = 12.8. In the reaction time task, θ is varied for each value of ¯β to find the maximal possible reward rate RR (see text); Tdel = 3 seconds. In the controlled duration case, θ = 15.3 is fixed, in agreement with a value indicated by behavioral data (see below and Figure 14). optimality under this definition as they tend to favor accuracy over speed in two-alternative forced choice trials (Zacksenhouse et al., 2010). Here, we simply use this quantity to formulate a scalar performance metric that provides a clear, compact interpretation of reaction time data. Plotted in Figure 11(A) are multiple accuracy vs. speed curves. The heavy solid line corresponds to the "baseline" model with robustness and mistuning set to zero (see Figure 4). The lighter solid line corresponds to the "mistuned" model with σβ = .1. The remaining dashed lines correspond to the "recovery" model for three different, nonzero levels of the robustness limit R. Also plotted in the background as dashed lines are RR isoclines -- that is, lines along which RR takes a constant value, with Tdel = 3 sec. On each accuracy vs. speed curve, there exists a RR-maximizing (RT, accuracy) pair. This corresponds to a tangency with one RR isocline, and is plotted as a filled circle. In general, each model achieves maximal RR via a different threshold θ; values are specified in the legend of Panel (B). (A general treatment of RR-maximizing thresholds for drift-diffusion models is given in Bogacz et al. (2006).) In sum, we see that mistuned integrators with a range of increasing robustness limits R achieve greater RR, as long as their thresholds are adjusted in concert. The optimal values of RR for a range of robustness limits R are plotted in Figure 11(B). This figure illustrates the fundamental tradeoff between robustness and sensitivity discussed above. If there is variability in feedback mistuning (σβ > 0), increasing R can help recover performance. However, beyond at a certain point increasing R further starts to diminish performance, as too much of the input signal is ignored. 3.4 Biased mistuning towards leak or excitation We next consider the possibility that variation in mistuning from trial to trial could occur with a systematic bias in favor of either leak or excitation, and ask whether the robustness limit has quali- tatively similar effects on decision performance as for the unbiased case studied above. Specifically, we draw the mistuning parameter β from a gaussian distribution with standard deviation σβ = 0.1 as above, but with various mean values ¯β (see Methods). In Figure 12(A) we show reward rates as a function of the bias ¯β, for several different levels of the robustness limit R. At each value of ¯β, the highest reward rate is achieved for a value of R > 0; that is, regardless of the mistuning bias, there 20 Accuracy RR Figure 13: Effect of the robustness limit R on decision performance in a controlled duration task, under the bounded integration model of Kiani et al (2008) (see text). Dot coherence C = 12.8. (A) Increasing the robustness limit R helps recover performance lost to mistuning at multiple reaction times in the controlled duration task. Specifically, moving from the "baseline" model to the "mistuned" model decreases decision accuracy (solid arrow), but this lost accuracy can be partially or fully recovered for R > 0 (dotted arrow). (B) When allowing for biased mistuning ( ¯β 6= 0, σ = .1), R still allows for recovery of performance; effects are most pronounced when ¯ β > 0. exists a R > 0 that will improve performance vs. the non-robust case ( R = 0). We note that this improvement appears minimal for substantially negative mistuning biases, but is more noticable for the values of ¯β that yield the highest RR. Finally, the ordering of the curves in Figure 12(A) shows that, for many values of ¯β, this optimal robustness limit is an intermediate value less than R = 2. While Figure 12 only assesses performance via a particular performance metric (RR, Tdel = 3 sec.), the analysis in "Reward rate and the robustness-sensitivity tradeoff" suggests that the result will hold for other performance metrics as well. Moreover, Figure 12(B) demonstrates the analogous effect for the controlled duration task: for each mistuning bias ¯β, decision accuracy increases over the range of robustness limits shown. 3.5 Bounded integration as a model of the fixed duration task We have demonstrated that increasing the robustness limit R can improve performance for mistuned integrators, in both the reaction time and controlled duration tasks. In the latter, a decision is made by examining which integrator had accumulated more evidence at the end of the time interval. In contrast, Kiani et al. (2008) argue that decisions in the controlled duration task are actually made with a decision threshold (or bound). That is, evidence accumulates toward a bound as in the reaction time task; if accumulated evidence crosses the bound before the end of the task duration, the subject simply waits for the opportunity to report the choice, ignoring any further evidence. Figure 13 demonstrates that our observations about the how the robustness limit can recover performance lost to mistuned feedback carry over to this model of decision making as well. Specif- ically, Panel 13(A) shows how setting R > 0 improves performance in a mistuned integrator. In fact, more of the lost performance (up to 100%) is recovered than in the previous model of the controlled duration task (cf. Figure 6(A)). Panel 13(B) extends this result to show that some value of R > 0 will recovers lost performance over a wide range of mistuning biases ¯β (cf. Figure 12(B)). 21 Accuracy Accuracy Controlled Duration (sec) 3.6 Compatibility of the robust integrator model with behavioral data Given the fact that the robustness property can improve decision performance in our model, we next ask whether robust limits R > 0 are compatible with known behavioral data. To answer this question, we fit accuracy and chronometric functions from robust integrator models to behavioral data of Roitman and Shadlen (2002) in the reaction time task. This fit is via least squares across the range of coherence values, and requires two free parameters: additive noise variance νγ (see Methods) and the decision bound θ. Figure 14 shows the results. Panels (A) and (B) display accuracy and chronometric data (dots) together with fits for various integrator models. First, the solid line gives the fit for the "baseline" model. The close match between model and data agrees with findings of prior studies (Mazurek et al., 2003). Next, the dashed and dotted lines give fits for mistuned models (σβ = 0.1), with three values of bias in feedback mistuning ( ¯β). To obtain these fits, both νγ and θ are changed from their values for the baseline case. In particular, the noise variance νγ is lowered when feedback is mistuned. This makes intuitive sense: we have seen in Figure 5 that mistuned feedback worsens performance for a given signal, so that matching a fixed dataset with a mistuned integrator requires improving the fidelity of the incoming signal. Figures 14(C), (D) show analogous results for robust integrators. For all cases in these panels, we take the robustness limit R = 1.15. We fix levels of additive noise to values found for the non-robust case above, on order to demonstrate that by adjusting the decision threshold, one can obtain approximate fits to the same data. This is expected from our results above: Figure 6 shows that, while accuracies at given reaction times are higher for mistuned robust vs. non-robust models, the effect is modest on the scale of the full range of values traced over an accuracy curve. Moreover, for the perfectly tuned case, accuracies at given reaction times are very similar for robust and non-robust integrators (Figure 7, with a slightly lower value of R). Thus, comparable pairs of accuracy and RT values are achieved for robust and non-robust models, leading to similar matches with data. In sum, the accuracy and chronometric functions in Figure 14 show that all of the models schematized in Figure 4 -- "baseline", "mistuned", "robust", and "recovery" -- are generally compatible with the chronometric and accuracy functions reported in Roitman and Shadlen (2002). In order to further test whether empirical data are consistent with the robust integrator model, we compared simulated reaction time histograms with those found in Roitman and Shadlen (2002). First, Figure 15 compares the reaction time histograms resulting from the "baseline" model (Panel A) and the "recovery" model (Panel B). These are plotted in blue; the red histograms are data are taken from Subject "B". In both panels, the histograms have similar means, but differ in their shape; in particular, the model predicts a broader range of reaction times and a more slowly decaying tail of the RT distribution. From these data, we conclude that neither the "baseline" nor the "recovery" model quantitatively reproduce the details of reaction time distributions, when the free parameters θ and σβ are constrained by fitting the accuracy and chronometric functions. An urgency signal was introduced in Churchland et al. (2008) to better capture behavioral and physiological data. We next incorporated such a signal into our model to determine whether it would better align our predicted reaction time histograms with the empirical data. We chose to implement urgency by assuming a collapsing decision bound, which decreases monotonically from a peak value of θ0 to a steady state value θss with a halflife t1/2: θ(t) = θ0 − (θ0 − θss) t t + t 1 2 . (23) 22 (A) (C) (B) (D) Perfect Tuning Mistuning β ∼ δ(0) ∼ N(cid:0)0, .12(cid:1) ∼ N(cid:0)−.05, .12(cid:1) ∼ N(cid:0).05, .12(cid:1) -- - - -.- ... Not Robust ( R = 0) Robust ( R = 1.15) (cid:0)θ, √νγ(cid:1) (15.3, 14.0) (11.0, 12.0) (9.2, 12.1) (13.8, 11.5) (7.6, 14.0) (6.8, 12.0) (5.7, 12.1) (8.6, 11.5) Figure 14: Accuracy (A,C) and chronometric (B,D) functions: data and model predictions. Solid dots are behavioral data for rhesus monkeys performing the dot-motion discrimination task (Roit- man and Shadlen, 2002). In each panel, the accuracy and chronometric functions are identified with behavioral data via a least-squares fitting procedure over the free parameters θ and νγ. In Panels (A,B), the robustness threshold R = 0, and results are shown for "baseline" and exemplar "mistuned" models (see legend in table). In Panels (C,D), R = 1.15, and results are shown for "robust" and "recovery" models. The close matches to data points indicate that each model can be broadly reconciled with known psychophysics. Parameter values for each curve are summarized in the included table. 23 Dot Coherence (C) Dot Coherence (C) Reaction Time (ms) RT (ms.) Accuracy Dot Coherence (C) Dot Coherence (C) Reaction Time (ms) RT (ms.) Accuracy Figure 15: Reaction time histograms, with decision bounds held constant in time and dot coherence C=12.8. Histograms for reaction times for our model with both R = 0 (A) and R = 1.15 (B) are plotted in blue. Overlayed are the reaction time histograms for one subject (Subject "B") in (Roitman and Shadlen, 2002), in red (semitransparent). Box-and-whisker plots indicate the quartiles for each data set. Both histograms have identical means, but clearly differ in basic shape (i.e., the model produces longer tails). We emphasize that this mismatch is a property of our basic integration to bound model, regardless of the value of the robustness limit R (see text). Figure 16: Reaction time histograms, with collapsing decision bounds. This shows that introducing collapsing bounds produces simulated RT histograms that are a closer match to data, for both non- robust (Panel A, R = 0) and robust (Panel B, R = 1.15) integrators. This shows that RT histograms can be similarly well matched to data for both cases. Here t 1 2 = 500, θ0 = 25, and θss = 0. 24 Reaction Time (ms) Reaction Time (ms) P[DT] P[DT] Reaction Time (ms) Reaction Time (ms) P[DT] P[DT] Figure 16 compares model reaction time histograms produced with the collapsing bound against the data, and indeed finds a closer fit: qualitatively, the improvement in fit is similar for both the non- robust ( R = 0) and robust ( R = 1.15) cases. In sum, this shows that the robust integrator model is capable of producing roughly similar patterns of reaction times compared with those observed experimentally. 4 Discussion A wide range of cognitive functions require the brain to process information over time scales that are at least an order of magnitude greater than values supported by membrane time constants, synaptic integration, and the like. Integration of evidence in time, as occurs in simple perceptual decisions, is one such well studied example, whereby evidence bearing on one or another alternative is gradually accumulated over time. This is formally modeled as a bounded random walk or drift- diffusion process in which the state (or decision) variable is the accumulated evidence for one choice and against the alternative(s). Such formal models explain both the speed and accuracy of a variety of decision-making tasks studied in both humans and nonhuman primates (Ratcliff, 1978; Luce, 1986; Gold and Shadlen, 2007; Palmer et al., 2005), and neural correlates have been identified in the firing rates of neurons in the parietal and prefrontal association cortex (Mazurek et al., 2003; Gold and Shadlen, 2007; Churchland et al., 2008; Shadlen and Newsome, 1996; Schall, 2001; Shadlen and Newsome, 2001; Kim et al., 2008). The obvious implication is that neurons must somehow integrate evidence supplied by the visual cortex, but there is mystery as to how. The reason this is a challenging problem is that the biological building blocks operate on rel- atively short time scales. From a broad perspective, the challenge is to assemble neural circuits that that can sustain a stable level of activity (i.e., firing rate) and yet retain the capability to increase or decrease firing rate when perturbed with new input (e.g., momentary evidence). A well known solution is to suppose that recurrent excitation might balance perfectly the decay modes of membranes and synapses (Cannon et al., 1983; Usher and McClelland, 2001). However, this balance must be fine tuned (Seung, 1996; Seung et al., 2000), or else the signal will either dis- sipate or grow exponentially (Figure (A), top). Several investigators have proposed biologically plausible mechanisms that mitigate somewhat the need for such fine tuning (Lisman et al., 1998; Goldman et al., 2003; Goldman, 2009; Romo et al., 2003; Miller and Wang, 2006; Koulakov et al., 2002). These are important theoretical advances because they link basic neural mechanism to an important element of cognition and thus provide grist for experiment. Although they differ in important details, many of the proposed mechanisms can be depicted as if operating on a scalloped energy landscape with relatively stable (low energy) values, which are robust to noise and mistuning in that they require some activation energy to move the system to a larger or smaller value (Figure (A), bottom; cf. (Pouget and Latham, 2002)). The energy landscape is a convenient way to view such mechanisms -- which we refer to as robust integrators -- because it also draws attention to a potential cost. The very same effect that renders a location on the landscape stable also implies that the mechanism must ignore information in the incoming signal (i.e., evidence). Here, we have attempted to quantify the costs inherent in this loss. How much loss is tolerable before the circuit misses substantial information in the input? How much loss is consistent with known behavior and physiology? We focused our analyses on a particular well-studied task because it offers critical benchmarks to assess both the potential costs of robustness to behavior and a gauge of the degree of robustness that might be required to mimic neurophysiological recordings with neural network models. Moreover, we know key statistical properties of the signal and noise to be accumulated over time, based on 25 firing properties in area MT. Our central finding is that ignoring a surprisingly large part of the motion evidence would have almost negligible impact on performance. Indeed, we found that speed and accuracy are preserved even when almost a full standard deviation of the input distribution is ignored. We also found that a similar degree of robustness provides protection of performance against mistuning of recurrent excitation. Although in general this protection is only partial (Figure 6), for the controlled duration task it can be nearly complete (Figure 13(A), T > 2) depending on the presence of a decision bound. We can appreciate the impact of robust integration intuitively by considering the distribution of random values that would increment the stochastic process of integrated evidence. Instead of imagining a scalloped energy surface, we simply replace all the small perturbations in integrated evidence with zeros. Put simply, if a standard integrator would undergo a small step in the positive or negative direction, a robust integrator instead stays exactly where it was. In the setting of drift-diffusion, this is like removing a portion of the distribution of momentary evidence (the part that lies symmetrically about zero) and replacing the mass with a delta function at 0. At first glance this appears to be a dramatic effect -- see the illustration of the distributions in Figure 8 -- and it is surprising that it would not result in strong changes in accuracy or reaction time or both. Three factors appear to mitigate this loss of momentary evidence. First, we showed that setting weak values of the input signal to zero can reduce both its mean and its standard deviation by a similar amount, creating compensatory effects that result in a small change to the input signal- to-noise ratio. Second, we showed that, surprisingly, the small loss of signal to noise that does occur would not result in any loss of accuracy if the accumulation were to the same bound as for a standard integrator. The cost would be to decision time, but mainly in the regime that is dominated by drift -- that is, the shorter decision times -- hence not a large cost overall. Third, even this slowing is mitigated by the temporal dynamics of the input. Unlike for idealized drift diffusion processes, real input streams possess finite temporal correlation. Left unchecked, this would imply greater variability in the integrated signal. Interestingly, removing the weakest momentary inputs reduces the temporal correlation of the noise component of the input stream. This can be thought of as allowing more independent samples in a given time period, thereby improving accuracy at a given response time. Our robust integrator framework shares features with existing models in sensory discrimination. The interval of uncertainty model of Smith and Vickers (1989) and the gating model of Purcell et al. (2010) ignore part of the incoming evidence stream, yet they can explain both behavioral and neural data. We suspect that the analyses developed here might also reveal favorable properties of these models. Notably, some early theories of signal detection also featured a threshold, below which weaker inputs fail to be registered -- so called high threshold theory (reviewed in (Swets, 1961)). The primary difference in the current work is to consider single decisions made based on an accumulation of many such thresholded samples (or a continuous stream of them). Although they are presented at a general level, our analyses make testable predictions. For example, they predict that pulses of motion evidence added to random dot stimulus would affect decisions in a nonlinear fashion consistent with a soft threshold. Such pulses are known to affect decisions in a manner consistent with bounded drift diffusion (Huk and Shadlen, 2005) and its implementation in a recurrent network (Wong et al., 2007). A robust integration mechanism further predicts that brief, stronger pulses will have greater impact on decision accuracy than longer, weaker pulses containing the same total evidence. However, we believe that the most exciting application of our findings will be to cases in which the strength of evidence changes over time, as expected in almost any natural setting. One simple example is for task stimuli that have an unpredictable onset time, and whose onset is not immediately obvious. For example, in the moving dots task, this would correspond to subtle 26 increases in coherence from a baseline of zero coherence. Our preliminary calculations agree with intuition that robust integrator mechanism will improve performance: in the period before the onset of coherence, less baseline noise would be accumulated; after the onset of coherence, the present results suggest that inputs will be processed with minimal loss to decision performance -- despite the continued ignoring of weak components. This intuition can be generalized to apply to a number of settings with non-stationary sensory streams. Many cognitive functions evolve over time scales that are much longer than the perceptual decisions we consider in this paper. Although we have focused on neural integration, it seems likely that many other neural mechanisms are also prone to drift and instability. Hence, the need for robustness may be more general. Yet, it is difficult to see how any mechanism can achieve robustness without ignoring information. If so, our finding may provide some optimism. Although we would not propose that ignorance is bliss, in the right measure it may be less costly than one would expect. References Abbott LF, Dayan P (1999) The effect of correlated variability on the accuracy of a population code. Neural Computation 11:91 -- 101. Bair W, Zohary E, Newsome WT (2001) Correlated firing in macaque visual area MT: time scales and relationship to behavior. Journal of Neuroscience 21:1676. Billingsley P (1986) Probability and measure Wiley-Interscience. Blackwell H (1953) Psychophysical thresholds; experimental studies of methods of measurement. Engineering Research Institute, Univ. of Michigan. Bogacz R, Brown E, Moehlis J, Holmes P, Cohen JD (2006) The physics of optimal decision making: A formal analysis of models of performance in two-alternative forced-choice tasks. Psy- chological Review 113:700 -- 765. Britten KH, Shadlen MN, Newsome WT, Movshon JA (1992) The analysis of visual mo- tion: a comparison of neuronal and psychophysical performance. The Journal of Neuro- science 12:4745 -- 4765. Britten KH, Shadlen MN, Newsome WT, Movshon JA (1993) Responses of neurons in macaque MT to stochastic motion signals. Visual Neuroscience 10:1157 -- 1169. Brown E, Gao J, Holmes P, Bogacz R, Gilzenrat M, Cohen JD (2005) Simple neural networks that optimize decisions. International Journal of Bifurcation Chaos in Applied Sciences and Engineering 15:803 -- 826. Brown E, Holmes P (2001) Modelling a simple choice task: Stochastic dynamics of mutually inhibitory neural groups. Stochastics and Dynamics 1:159 -- 191. Burden R, Faires J (2011) Numerical analysis Brooks Cole. Cannon SC, Robinson DA, Shamma S (1983) A proposed neural network for the integrator of the oculomotor system. Biological Cybernetics 49:127 -- 136. Churchland AK, Kiani R, Shadlen MN (2008) Decision-making with multiple alternatives. Nature Neuroscience 11:693 -- 702. 27 Cohen MR, Newsome WT (2009) Estimates of the contribution of single neurons to perception depend on timescale and noise correlation. Journal of Neuroscience 29:6635 -- 6648. Donner TH, Siegel M, Fries P, Engel AK (2009) Buildup of Choice-Predictive Activity in Human Motor Cortex during Perceptual Decision Making. Current Biology 19:1581 -- 1585. Doya K, editor (2007) Bayesian brain: probabilistic approaches to neural coding The MIT Press. Edwards W (1965) Optimal strategies for seeking information: Models for statistics, choice reac- tion times, and human information processing. Journal of Mathematical Psychology 2:312 -- 329. Fuchs AF (1967) Saccadic and smooth pursuit eye movements in the monkey. The Journal of Physiology 191:609 -- 631. Gardiner C (2002) Handbook of stochastic methods: for physics, chemistry and the natural sciences Springer. Gillespie DT (1996) Exact numerical simulation of the Ornstein-Uhlenbeck process and its integral. Physical Review E 54:2084 -- 2091. Gold JI, Shadlen MN (2002) Banburismus and the Brain Decoding the Relationship between Sensory Stimuli, Decisions, and Reward. Neuron 36:299 -- 308. Gold JI, Shadlen MN (2007) The neural basis of decision making. Annual Review of Neuro- science 30:535 -- 574. Goldman MS (2009) Memory without Feedback in a Neural Network. Neuron 61:621 -- 634. Goldman MS, Levine JH, Major G, Tank DW, Seung HS (2003) Robust Persistent Neural Activity in a Model Integrator with Multiple Hysteretic Dendrites per Neuron. Cerebral Cor- tex 13:1185 -- 1195. Green D, Swets J (1966) Signal detection theory and psychophysics Peninsula Pub. Hesse CH (1991) The one-sided barrier problem for an integrated ornstein-uhlenbeck process. Stochastic Models 7:447 -- 480. Higham DJ (2001) An algorithmic introduction to numerical simulation of stochastic differential equations. SIAM Review 43:525 -- 546. Huk AC, Shadlen MN (2005) Neural Activity in Macaque Parietal Cortex Reflects Temporal Integration of Visual Motion Signals during Perceptual Decision Making. Journal of Neuro- science 25:10420 -- 10436. Kemeny J, Snell J (1960) Finite markov chains D. Van Nostrand. Kiani R, Hanks TD, Shadlen MN (2008) Bounded integration in parietal cortex underlies de- cisions even when viewing duration is dictated by the environment. The Journal of Neuro- science 28:3017 -- 3029. Kim S, Hwang J, Lee D (2008) Prefrontal Coding of Temporally Discounted Values during In- tertemporal Choice. Neuron 59:161 -- 172. Koulakov AA, Raghavachari S, Kepecs A, Lisman JE (2002) Model for a robust neural integrator. Nature Neuroscience 5:775 -- 782. 28 Laming D (1968) Information theory of choice-reaction times Academic Press New York. Link SW, Heath RA (1975) A sequential theory of psychological discrimination. Psychome- trika 40:77 -- 105. Lisman J, Fellous J, Wang XJ (1998) A role for NMDA-receptor channels in working memory. Nature Neuroscience . Luce R (1963) A Threshold Theory for Simple Detection Experiments. Psychological Re- view 70:61 -- 79. Luce R (1986) Response times: their role in inferring elementary mental organization Oxford University Press, Oxford psychology series, no. 8. Machens CK, Romo R, Brody CD (2005) Flexible control of mutual inhibition: A neural model of two-interval discrimination. Science 307:1121 -- 1124. Mazurek ME, Roitman JD, Ditterich J, Shadlen MN (2003) A Role for Neural Integrators in Perceptual Decision Making. Cerebral Cortex 13:1257 -- 1269. Mazurek ME, Shadlen MN (2002) Limits to the temporal fidelity of cortical spike rate signals. Nature Neuroscience 5:463 -- 471. McKoon G, Ratcliff R (2008) The diffusion decision model: Theory and data for two-choice decision tasks. Neural Computation 20:873 -- 922. Miller P (2006) Analysis of spike statistics in neuronal systems with continuous attractors or multiple, discrete attractor states. Neural Computation 18:1268 -- 1317. Miller P, Wang XJ (2006) Power-law neuronal fluctuations in a recurrent network model of parametric working memory. Journal of Neurophysiology 95:1099 -- 1114. Nikitchenko M, Koulakov A (2008) Neural integrator: A sandpile model. Neural Computa- tion 20:2379 -- 2417. Okamoto H, Fukai T (2009) Recurrent network models for perfect temporal integration of fluctu- ating correlated inputs. PLoS Computational Biology 5:e1000404. Palmer J, Huk AC, Shadlen MN (2005) The effect of stimulus strength on the speed and accuracy of a perceptual decision. Journal of Vision 5:376 -- 404. Pouget A, Latham P (2002) Digitized neural networks: long-term stability from forgetful neurons. Nature Neuroscience 5:709 -- 710. Purcell BA, Heitz RP, Cohen JY, Schall JD, Logan GD, Palmeri TJ (2010) Neurally constrained modeling of perceptual decision making. Psychological Review 117:1113 -- 1143. Ratcliff R (1978) A Theory of Memory Retrieval. Psychological Review 85:59 -- 108. Ratcliff R, Rounder JN (1998) Modeling response times for two-choice decisions. Psychological Science 9:347 -- 356. Roitman JD, Shadlen MN (2002) Response of neurons in the lateral intraparietal area during a combined visual discrimination reaction time task. Journal of Neuroscience 22:9475 -- 9489. 29 Romo R, Kepecs A, Brody CD (2003) Basic mechanisms for graded persistent activity: discrete attractors, continuous attractors, and dynamic representations. Current Opinion In Neurobiol- ogy 13:204 -- 211. Salzman CD, Murasugi CM, Britten KH, Newsome WT (1992) Microstimulation in visual area MT: effects on direction discrimination performance. The Journal of Neuroscience 12:2331 -- 2355. Schall JD (2001) Neural basis of deciding, choosing and acting. Nature Reviews Neuro- science 2:33 -- 42. Seung HS (1996) How the brain keeps the eyes still. Proceedings of the National Academy of Sciences 93:13339 -- 13344. Seung HS, Lee DD, Reis BY, Tank DW (2000) Stability of the memory of eye position in a recurrent network of conductance-based model neurons. Neuron 26:259 -- 271. Shadlen MN, Britten KH, Newsome WT, Movshon JA (1996) A computational analysis of the relationship between neuronal and behavioral responses to visual motion. The Journal of Neuro- science 16:1486 -- 1510. Shadlen MN, Newsome WT (1996) Motion perception: seeing and deciding. Proceedings of the National Academy of Sciences 93:628 -- 633. Shadlen MN, Newsome WT (2001) Neural basis of a perceptual decision in the parietal cortex (area LIP) of the rhesus monkey. Journal of Neurophysiology 86:1916 -- 1936. Smith P (2010) From Poisson shot noise to the integrated Ornstein-Uhlenbeck process: Neurally principled models of information accumulation in decision-making and response time. Journal of Mathematical Psychology 54:266 -- 283. Smith PL, Ratcliff R (2004) Psychology and neurobiology of simple decisions. TRENDS in Neurosciences 27:161 -- 168. Smith PL, Vickers D (1989) Modeling evidence accumulation with partial loss in expanded judg- ment. Journal of Experimental Psychology: Human Perception and Performance 15:797. Swets JA (1961) Is there a sensory threshold. Science 134:168 -- 177. Usher M, McClelland JL (2001) The time course of perceptual choice: the leaky, competing accumulator model. Psychological Review 108:550 -- 592. Wald A (1944) On cumulative sums of random variables. Annals Of Mathematical Statis- tics 15:342 -- 342. Wald A (1945) Sequential Tests of Statistical Hypotheses. The Annals of Mathematical Statis- tics 16:117 -- 186. Wald A, Wolfowitz J (1948) Optimum character of the sequential probability ratio test. The Annals of Mathematical Statistics 19:326 -- 339. Wang XJ (2002) Probabilistic decision making by slow reverberation in cortical circuits. Neu- ron 36:955 -- 968. Wang XJ (2008) Decision Making in Recurrent Neuronal Circuits. Neuron 60:215 -- 234. 30 Wong K, Huk A, Shadlen M, Wang XJ (2007) Neural circuit dynamics underlying accumula- tion of time-varying evidence during perceptual decision making. Frontiers in Computational Neuroscience 1. Wong KF, Wang XJ (2006) A recurrent network mechanism of time integration in perceptual decisions. Journal of Neuroscience 26:1314 -- 1328. Zacksenhouse M, Bogacz R, Holmes P (2010) Robust versus optimal strategies for two-alternative forced choice tasks. Journal of Mathematical Psychology 54:230 -- 246. Zohary E, Shadlen MN, Newsome WT (1994) Correlated neuronal discharge rate and its impli- cations for psychophysical performance. Nature 370:140 -- 143. 31 Title: Supplementary Materials for Neural integrators for decision making: A favorable tradeoff between robustness and sensitivity Abbreviated Title: Supplementary Materials for Neural integrators for decision making Authors: Nicholas Cain, Department of Applied Mathematics, University of Washington Andrea K. Barreiro, Department of Applied Mathematics, University of Washington Michael Shadlen, Department of Physiology and Biophysics, University of Washington Eric Shea-Brown, Department of Applied Mathematics, University of Washington Corresponding Author: Eric Shea-Brown Department of Applied Mathematics Box 325420 Seattle, WA 98195-2420 [email protected] Number of pages: 10 Number of figures: 7 Number of tables: 0 Conflict of Interest: None Acknowledgements: This research was supported by a Career Award at the Scientific Interface from the Burroughs-Wellcome Fund (ESB), the Howard Hughes Medical Institute, the National Eye Institute Grant EY11378, and the National Center for Research Resources Grant RR00166 (MS), by a seed grant from the Northwest Center for Neural Engineering (ESB and MS), and by NSF Teragrid allocation TG-IBN090004. arXiv:1111.6573v1 [q-bio.NC] 28 Nov 2011 1 1 Derivation of effective model Here we we describe a family of neural circuit models, parameterized by a robustness value R, that produce dynamics similar to that of the central robust integrator model of the main paper (given by Equation 5 in the main text). In particular, as R → 0, the dynamics reduce to a perfect integrator. Our construction follows closely that of Koulakov et al. (2002) and Goldman et al. (2003). The circuits that we will study derive their robustness from multistability, which follows from recurrent self-excitation among multiple subpopulations (or subunits). We begin with a system of N differential equations, one for each subunit. Depending on the activity in the rest of the network, each individual subunit can be become bistable, so that its eventual steady-state value (i.e., "On" or "Off") depends on its past. This effect, known as hysteresis, underlies several robust integrator models (Koulakov et al., 2002; Goldman et al., 2003). The circuit integrates inputs via sequential activation of subunits, in an order determined by graded levels of "background" inputs (or biases) to each subunit. Following Goldman et al. (2003), we collapse the N differential equations that describe individual subunits into a equation that approximates the dynamics of the entire integrator. This expression for the total firing rate E(t) averaged over all subpopulations reduces to the robust integrator model of the main text (Equation 5 in the main text). 1.1 Firing rate model The firing rate ri(t) of the ith bistable subunit (i ∈ {1, 2, ..., N}) is modeled by a firing rate equation: τE dri dt = −ri + f (ri) (1) N X i6=j rj + a∆I − bi f (ri) = r− + (r+ − r−)Hp ∗ ri + q(1 + β) The parameters and variables in this equation are as follows: • ri(t): Firing rate of subunit (or pool) • a∆I(t) : Input signal, with synaptic weight • p: Within-pool "synaptic" weight • q: Between-pool "synaptic" weight • β: Fractional mistuning of feedback connectivity • N : Number of subunits in the integrator network • r−: Minimum firing rate of a subunit • r+: Maximum firing rate of a subunit • τE : Time constant governing firing rate dynamics We also define: • E(t) = 1 NPN i=1 ri(t): Overall integrator activity 2 (A) (B) Figure 1: Simultaneous plots of the identity line and the "feedback line," G(E) for two circuits with different numbers of subunits. (A) Here N = 1, and so the feedback line G(E) = f (E) is exactly determined as a function of E = r1 (see Equation (3)). The two intersections c0 and c1 are stable fixed points. In this way, the subunit's firing rate is bistable, and the value attained will depend on the history of the circuit activity. As ∆I is changed, this translates f (E), eventually eliminating either co or c1 and forcing the subunit to the remaining stable fixed point (here, c0 corresponds to E = 8 Hz. and c1 to 40 Hz). In dashed curves, the feedback line is plotted for two such values of ∆I. (B) Now N > 1 and so the feedback line G(E) is no longer unambiguously specified as a function of its argument. The function is instead the sum of N potentially bi-valued functions, who's actual values will depend on the stimulus history. We represent this fact by plotting the feedback line as a set of stacked boxes, representing the potential contribution of the ith subunit to the total integrator dynamics (Goldman et al., 2003). Figure 1(A) demonstrates the firing rate dynamics of a circuit composed of a single subunit (N = 1). Here we refer to Equation (1) and plot two curves vs. E = r1. The first is the identity line, corresponding to the first "decay" term in Equation (1). The second is the "feedback" line, defined by the second term in Equation (1). Since N = 1, this simplifies to f (r1). The two intersections marked c0 and c1 are stable fixed points (which we refer to as "On" and "Off" respectively). Thus, the subunit shown is bistable. Importantly, however, the location of the step in f (E) varies with changes in the input signal (as per Equation 1). In particular, substantial values of ∆I(t) will (perhaps transiently) eliminate one of the fixed points, forcing the subunit into either the "On" or the "Off" state with ri = r+ or ri = r−, respectively. Moreover, the change is self-reinforcing via the recurrent excitation pri. The range over which a given subunit displays bistability is affected by the mistuning parameter β, which scales the total recurrent excitation from the rest of the circuit. We now derive the dynamics of the overall firing rate E(t). After summing both sides of Equation (1) over i we have: where G(E) = r− + (r+ − r−) N N τE dE dt = −E + G(E) H [(p − q(1 + β))ri + N q(1 + β)E + a∆I − bi] . (2) (3) X i=1 At this point, we almost have a differential equation for a single variable, E(t). However, Equa- tion (3) still depends on the N activities ri of the individual subunits, and at any particular time their values are not uniquely determined by the value of E; we can only bound their values as r− ≤ ri ≤ r+. We will return to discuss the dynamics of Equation 2 below. 3 (A) (B) Figure 2: Plot of possible equilibria for Equation 2. (A) The extent of each multivalued feedback function defines the minimum input necessary to perturb the system away from equilibrium, defining the "fixation" lines. As N → ∞, the stable fixed points become more tightly packed on the interval (r−, r+). (B) When the integrator is mistuned, the fixation lines and the feedback line are no longer parallel. The rate that the integrator accumulates input is approximated by the distance between the center line of the feedback subunits ( G(E)), and the feedback line. However, integration only occurs when the "fixation condition" is no longer satisfied, i.e. when the feedback line is no longer bounded by the fixation lines. 1.2 Bias term The bias term for the ith subunit, bi, is set by analyzing the range of values of E for which the exact value of the feedback function f (ri) is unknown. In the case of an integrator composed of only a single subunit, we choose the bias term to cause the positive input needed to force the unit to be on, and the negative input needed to force the unit to be off, to take the same values. This yields b1 = p(r++r−) . 2 The general case of N subunits is more complicated. Now the feedback contribution of the ith unit, f (ri), is no longer a simple function of the population activity E. Instead, it has additional dependence on its own activity ri. We see this clearly in Equation 3, where the values of ri that contribute to the definition of G(E) are unspecified. However, we do know that each ri is trapped between r+ and r−. Therefore, we can plot G(E) as the sum of a sequence of bivalued functions of E; see Figure 1(B) and (Goldman et al., 2003). The contribution from each pool is then represented by the shaded region in the plot. Finally, the bias terms are chosen to center these shaded boxes over the identity line. The correct biases in this general case are: bi = p(cid:18) r− + r+ 2 (cid:19) + q((i − 1)r+ + (N − i)r−) . (4) 1.3 Fixation lines We next define the "fixation" lines ( G+(E) and G−(E)), which are the consequence of this multi- valued property of the integrator. These lines define a region (the fixation region) that runs across the outermost corners of the "stacked boxes" in Figure 2. G+(E) = E(1 + β) + a∆I N q − βr+ N + p(r+ − r−) 2N q (5) 4 G−(E) = E(1 + β) + a∆I N q − βr− N − p(r+ − r−) 2N q . (6) The term fixation region refers to the following property: if the input ∆I is such that the identity line lies within the fixation region, then the integrator will possess a range of closely spaced fixed points (where E = G(E)). Thus, E is not expected to change from its current value; in other words, integration of ∆I will not occur (Goldman et al., 2003; Koulakov et al., 2002). Recall that ∆I acts to shift these boxes leftward or rightward relative to the identity line, just as in the analysis of Figure 1. As a consequence, it is weak inputs that fail to be integrated. From this analysis, we can see that integration by the system as a whole relies on two con- cepts. The first is a condition on ∆I necessary to eliminate fixed points; we call this the "fixation condition." The second is the question of how quickly to integrate once this condition is no longer satisfied. We address this next. 1.4 Integration Based on the analysis above, we derive a reduced model that approximately captures the dynamics of the "full" model indicated by Equation 2. We call this the "effective" model. The rate of change of E -- i.e., the rate of integration -- is given by the distance between the the current value of G(E) and the identity line. We approximate this by the distance between the middle of the fixation lines, which we define as G(E), and the identity line. This is pictured in Figure 2(B), and yields: τE dE dt = −E + G(E) ≈ −E + G(E) G(E) = (1 + β)E + a∆I N q − β (r+ + r−) 2N (7) (8) We emphasize that integration by this equation only occurs when the "fixation" condition is no longer satisfied, i.e. when the fixation lines no longer bound the the identity line. 1.5 Fixation condition The last step in defining the 1-dimensional "effective" model is determining the fixation condition. We must solve for the values of E that cause the feedback line to lie between the two fixation lines: No Change in E ⇐⇒ G−(E) < E < G+(E) ⇐⇒ G(E) − ⇐⇒ (cid:12)(cid:12)(cid:12)(cid:12)βE + (r+ − r−)(p − qβ) 2N q (cid:12) (cid:12) (r+ + r−) (cid:12) (cid:12) < a∆I N q − β 2N (r+ − r−)(p − βq) 2N q < E < G(E) + (r+ − r−)(p − qβ) 2N q (9) (10) (11) If this condition is violated with ∆I = 0, the integrator displays runaway integration (β > 0) or leak (β < 0). If it is satisfied when ∆I = 0, we have a condition on the level of ∆I that must be present for integration to occur. This yields a piecewise-defined differential equation, corresponding to when integration can and cannot occur: 5 0 βE + a∆I N q − β (r++r−) 2N :(cid:12)(cid:12)(cid:12)βE + a∆I : N q − β (r++r−) otherwise 2N (cid:12)(cid:12)(cid:12) < (r+−r−)(p−βq) 2N q τE dE dt ≈ 2N q We now simplify this equation as follows. First, we assume that p is much larger than βq, so 2N q = R, where R is the robustness parameter in the main text. Next, we becomes negligible (here the synaptic weights N q . This that (r+−r−)(p−βq) note that as N increases, the additive term β (r++r−) a and p can adjusted so that the remaining terms do not vanish). Finally, we define κ = a yields the central Equation 5 of the main text: ≈ (r+−r−)p 2N τE dE dt =(cid:26) 0 βE + κI : βE + κ∆I ≤ R : otherwise (12) 1.6 Quality of the model reduction Figure 3 compares the reduced "effective" model, defined by Equation 12 and denoted by the variable E, and the "full model" of multiple subunits, defined by Equation 1 and denoted by the variable E. Results are given for three different values of the normalized robustness limit R 1 are displayed, all in response to the same input signal ∆I(t) for ease of comparison. As R increases, the ability of the effective model to track the full model decreases. The quality of the reduction can be quantified by examining the relative error between the full and effective models: (t) = E(t) − E(t) E(t) (13) Histograms of (t) evaluated at t = 500 sec. are given in Figure 3(D-F). The agreement of the two models is within roughly 20% across a range of robustness values R. This is sufficient for our purposes of demonstrating an approximate connection between the simplified integrator model used in the main text and one of its many possible neural substrates. 2 Additional indicators that robust integration is compatible with empirical data from decision tasks As emphasized in the main text, we make several assumptions about how evidence is integrated over time in our model of perceptual decision making. The first is that the tuning of connectivity in the integrator circuit is inherently imprecise from trial to trial. This imprecision in feedback tuning results in spurious activity, in the form of either exponential growth or decay. To counter this, we consider the presence of a robustness mechanism meant to ameliorate this problem by allowing the integrator to fixate at various levels of activity despite mistuning. These assumptions differ from, e.g., the drift-diffusion model for two-alternative decision mak- ing, which has been shown to explain psychophysical and physiological data. An important question is therefore whether the model at hand will also be able to capture these data. In the main text we find that the answer is yes, for behavioral data on decision accuracy and speed (Figures 14 and 16 in the main text). We next expand this analysis by considering several other aspects of the data: 1This is the robustness parameter divided by the steady-state standard deviation of the input signal; see main text. 6 Figure 3: Quality of the reduction from the "full" model to the "effective" model. For three different values of the robustness limit R, traces E(t) for the full model (Equation 1), and its reduction E(t) (Equation 1.5) are compared (Panels A, B, and C). Here, the same signal is integrated by each model. Panels D, E, and F show histograms (N=100) of the relative error between the two models at t = 500 ms. (See Equation 13). The mean µ and standard deviation σ for each distribution are also indicated. At these levels of R, the effective model gives an approximation of the full subpopulation-based model with low to moderate error (see text). 7 Integrator Traces Relative Error t (ms) (A) (B) Figure 4: Reaction time histograms from Figures 15 and 16 of the main text, sorted by correct and incorrect responses and plotted in Panels (A) and (B), respectively. Correct trials are pictured above, while incorrect trials are below; settings as in the table included in Figure 14 of the main text. (A) Mean reaction time for error trials is 740 ms., vs. 671 ms. for correct trials; here a non-robust integrator is precisely tuned (σβ = 0, and R = 0) (B) For a robust integrator with imprecise tuning, error trials are now shorter (644 vs. 652 ms.). error-trial vs. correct-trial reaction time histograms, controlled duration accuracy performance, decision-triggered stimulus predictions, and traces of neural firing rates vs. time. 2.1 Asymmetry of reaction times between correct and incorrect trials Roitman and Shadlen (2002) and many other authors report differences in mean reaction times when responses are sorted by correct vs. incorrect responses. In Figure 4, we perform the same analysis on the data in Figure 14 of the main text. First, Panel (A) shows that our model of a perfectly tuned, non-robust integrator produces the shows the same qualitative asymmetry, with error trials on average having longer reaction times. With the addition of imprecise tuning and robustness, however this asymmetry can be eliminated (Panel (B)). Numerous changes to the model could recover the longer error RTs. One possibility is to introduce a bias in mistuning parameters: choosing ¯β > 0 has the desired effect. Others include inclusion of collapsing bounds, urgency signals, and other forms of trial-to-trial variability (for example Ratcliff and Rounder (1998)). 2.2 Model consistency with controlled duration data In the main text, we showed that an integrator model with imprecise tuning and robustness can reproduce the chronometric and psychometric curves reported in Roitman and Shadlen (2002) for the free-response task. We also considered the consequences of both robustness and mistuning in a controlled duration task. Here we show that the same parameters (Table, Figure 14 of the main text), this model can reproduce behavioral data reported in Kiani et al. (2008) for the controlled duration task. We assume bounded integration, where an integration threshold determines which alternative is selected, and the subject waits until the end of the trial to report the selection. The comparison is demonstrated in Figure 5, where model the accuracy functions from the model are overlaid on 8 500 1000 1500 RT (ms) 2000 P[Correct] P[Incorrect] 500 1000 1500 RT (ms) 2000 P[Correct] P[Incorrect] (A) (B) Figure 5: Comparison of model with controlled duration accuracy, C=12.8. The perfectly tuned model (A), and imprecise/robust model (B), both predict accuracy that is comparable with those reported in (Kiani et al., 2008) (Dotted lines, overlaid). In both cases, the CD task is modeled with bounded integration, with threshold set as in the table of Figure 14, main text. the data provided in the Kiani et al. study, with no parameter changes (including threshold) from the data reported in the fit to the Roitman et al. reaction time task data in the main text. 2.3 Decision-triggered stimuli Kiani et al. (2008) also report the time-course of motion evidence leading to the eventual decision in the controlled duration task. Specifically, the authors study long duration trials with neutral stimuli (i.e., zero coherence, so neither alternative is "correct"). In our model, motion evidence is encoded by the stimulus ∆I(t) (see main text). In Figure 6 we sort 400 trials (coherence C=0) based on which alternative was selected. We average the stimuli, again assuming bounded integration in a controlled duration task. We observe that the results of our model, both perfectly tuned and imprecisely tuned with nonzero robustness, are consistent with those reported by Kiani et al. Specifically, the early separa- tion of motion energy profiles for rightward (red) and leftward (blue) choices are indicative bounded accumulator, and do not qualitatively change when mistuning and robustness are included in the model. 2.4 Trial-sorted traces of firing rates vs. time Firing rates of LIP neurons reflect the time course of accumulated sensory evidence (Gold and Shadlen, 2007). We verified that our model displays analogous ramping activity when averaged over the similar number of trials. In particular, we checked this both for the perfectly tuned, non-robust version of our model (Figure 7(A-B)) and in the presence of mistuning and robustness (Figure 7(C-D)). Integrator activity averaged across multiple trials is averaged, and aligned to both the onset 9 T (sec) Accuracy T (sec) Accuracy (A) (B) Figure 6: Time course of decision-triggered stimuli. Stimulus realizations ∆I(t) are sorted by "left" vs "right" alternative selected (red vs. blue; neither alternative is "correct" as coherence C=0), and averaged across trials. Again assuming perfect tuning, (A), and imprecise tuning with robustness, (B), we find that both results are consistent with the motion energy profiles reported in Kiani et al. (2008). All integrator settings are the same as reported in the table in Figure 14 of the main text. (A) (C) (B) (D) Figure 7: Time course of model integrator activity E(t), during stimulus integration in reaction time tasks. Integration of input signal at various coherence values, with responses aligned at the beginning (A, C) and end (B, D), averaged across 400 trials. As expected, activity ramps up faster as dot coherence increases in both the perfectly tuned (A, B) and imprecisely tuned/robust (C, D) cases, in accordance with physiological observations. (For example, Roitman and Shadlen (2002).) 10 of stimulus and upon reaching threshold in both model setups. Activity ramps up (integrates) faster at higher dot coherence values, consistent with physiological measurements. This provides further evidence that our two additional model assumptions are consistent with known physiology for circuits involved in perceptual decision making. References Gold JI, Shadlen MN (2007) The neural basis of decision making. Annual Review of Neuro- science 30:535 -- 574. Goldman MS, Levine JH, Major G, Tank DW, Seung HS (2003) Robust Persistent Neural Activity in a Model Integrator with Multiple Hysteretic Dendrites per Neuron. Cerebral Cor- tex 13:1185 -- 1195. Kiani R, Hanks TD, Shadlen MN (2008) Bounded integration in parietal cortex underlies de- cisions even when viewing duration is dictated by the environment. The Journal of Neuro- science 28:3017 -- 3029. Koulakov AA, Raghavachari S, Kepecs A, Lisman JE (2002) Model for a robust neural integrator. Nature Neuroscience 5:775 -- 782. Ratcliff R, Rounder JN (1998) Modeling response times for two-choice decisions. Psychological Science . Roitman JD, Shadlen MN (2002) Response of neurons in the lateral intraparietal area during a combined visual discrimination reaction time task. Journal of Neuroscience 22:9475 -- 9489. 11
1606.00821
2
1606
2017-01-09T19:42:45
Towards a statistical mechanics of consciousness: maximization of number of connections is associated with conscious awareness
[ "q-bio.NC" ]
It has been said that complexity lies between order and disorder. In the case of brain activity, and physiology in general, complexity issues are being considered with increased emphasis. We sought to identify features of brain organization that are optimal for sensory processing, and that may guide the emergence of cognition and consciousness, by analysing neurophysiological recordings in conscious and unconscious states. We find a surprisingly simple result: normal wakeful states are characterised by the greatest number of possible configurations of interactions between brain networks, representing highest entropy values. Therefore, the information content is larger in the network associated to conscious states, suggesting that consciousness could be the result of an optimization of information processing. These findings encapsulate three main current theories of cognition, as discussed in the text, and more specifically the conceptualization of consciousness in terms of brain complexity. We hope our study represents the preliminary attempt at finding organising principles of brain function that will help to guide in a more formal sense inquiry into how consciousness arises from the organization of matter.
q-bio.NC
q-bio
Towards a statistical mechanics of consciousness: maximization of number of connections is associated with conscious awareness R. Guevara Erra1, D. M. Mateos2, R. Wennberg3, J.L. Perez Velazquez2∗ 1 Laboratoire Psychologie de la Perception, CNRS and Université Paris Descartes, Sorbonne Paris Cité, Paris, France. 2 Neuroscience and Mental Health Programme, Division of Neurology, Hospital for Sick Children. Institute of Medical Science and Department of Paediatrics, University of Toronto, Toronto, Canada. 3 Krembil Neuroscience Centre, Toronto Western Hospital, University of Toronto, Toronto, Canada. * [email protected] January 11, 2017 Abstract It has been said that complexity lies between order and disorder. In the case of brain activity, and physiology in general, complexity issues are being considered with increased emphasis. We sought to identify features of brain organization that are optimal for sensory processing, and that may guide the emergence of cognition and consciousness, by analysing neurophysiological recordings in conscious and unconscious states. We find a surprisingly simple result: normal wakeful states are characterised by the greatest number of possible configurations of interactions between brain networks, representing highest entropy values. Therefore, the information content is larger in the network associated to conscious states, suggesting that consciousness could be the result of an optimization of information processing. These findings encapsulate three main current theories of cognition, as discussed in the text, and more specifically the conceptualization of consciousness in terms of brain complexity. We hope our study represents the preliminary attempt at finding organising principles of brain function that will help to guide in a more formal sense inquiry into how consciousness arises from the organization of matter 1 Introduction How consciousness arises from the organization of matter is a subject of debate that spans several disciplines, from philosophy to physics. Multitude of studies focus on the investigation of patterns of synchrony in brain activity based on magnitudes of a variety of synchrony indices, and thus the search for organising principles of brain function is today more crucial than ever. We sought to identify global features of brain organization that are optimal for sensory processing and that may guide the emergence of conscious awareness. Our results provide a (very simple) answer to the question of what the magnitudes of synchrony indices represent in terms of the structure of brain activity. Neurophysiological recordings of brain activity demonstrate fluctuating patterns of cellular interactions, variability that allows for a wide range of states, or configurations of connections of distributed networks exchanging information, and support the flexibility needed to process sensory inputs. Recent years have seen a surge in the study of fluctuations in brain coordinated activity, studies that have raised conceptual frameworks 1 7 1 0 2 n a J 9 ] . C N o i b - q [ 2 v 1 2 8 0 0 . 6 0 6 1 : v i X r a such as that of metastable dynamics [1] and that have motivated interest in the practical application of assessments of nervous system variability for clinical purposes [2, 3]. A prominent question is how to describe the organizing principles of this collective activity, which allow features associated with consciousness to emerge. What is the optimal brain organization that allows it to adequately process sensory stimuli and enable the organism to adapt to its environment? Previous studies have revealed values of different indicators of brain coordinated activity, such as synchronization, associated with healthy and pathologic states by comparison of baseline values and those in, for instance, unconscious states like coma and epileptic seizures [4, 5, 6]. These observations prompt the question of what physiological organization underlies the specific values of the synchrony indices found in normal alert states and other conditions; in other words, what is special about these values found in conscious states? We believe that we have provided an answer to this question in our work. We propose that there is a certain general organization of brain cell ensembles that will be optimal for conscious awareness, that is, for the processing of sensory inputs. As an extension of previous work [7] where it was proposed that a general organising principle of natural phenomena is the tendency toward maximal – more probable – distribution of energy/matter, we venture that the brain organization optimal for conscious awareness will be a manifestation of the tendency towards a widespread distribution of energy (or, equivalently, maximal information exchange). Whereas we do not deal with energy or information in our work, we instead focus on the number of (micro)states, or combinations of connected signals derived from specific types of neurophysiologic recordings. We use the term "information" in the intuitive sense that normally permeates neuroscience: cell ensembles that are functionally connected to process/exchange information; furthermore, the equivalence between information exchange and energy transactions has been the subject of several studies, more specifically in [7] (see also [8, 9]). The question then becomes: how do we capture the nature of these organizations of cell interactions? We have followed the classic approach in physics when it comes to understanding collective be- haviours of systems composed of a myriad of units: the assessment of the number of possible config- urations, or microstates, that the system can adopt. In our study we focus on the collective level of description and assume that coordinated patterns of brain activity evolve due to interactions of meso- scopic areas ( [10, 11]). Thus we use several types of brain recordings in conscious and unconscious states, evaluating the number of "connections" between these areas and the associated entropy and complexity. We present evidence that conscious states result from higher entropy and complexity in the number of configurations of pairwise connections. The number of pairwise channel combinations is near the maximum of all possible configurations when the individual is processing sensory inputs in a normal manner (e.g. with open eyes). Our interpretation is that a greater number of configurations of interactions allows the brain to optimally process sensory information, fostering the necessary variabil- ity in brain activity needed to integrate and segregate sensorimotor patterns associated with conscious awareness. 2 Materials and Methods Electrophysiological recordings Recordings were analysed from 9 subjects, using magnetoencephalography (MEG), scalp electroen- cephalography (EEG) or intracranial EEG (iEEG). Three epilepsy patients were studied with MEG; 2 1 epilepsy patient was studied with iEEG; 3 epilepsy patients were studied with simultaneous iEEG and scalp EEG; and 2 nonepileptic subjects were studied with scalp EEG. For the study of seizures versus alert states, the 3 subjects with MEG recordings and the one with iEEG were used. Details of the patients' epilepsies and seizure types have been presented in previous studies (MEG patients in [5]; iEEG patients in [12]). For the study of sleep versus alert states, the 3 patients with combined iEEG and scalp EEG have been described previously (patients 1, 3, 4 in [13]); the 2 subjects studied with scalp EEG alone had been investigated because of a suspected history of epilepsy, but both were ulti- mately diagnosed with syncope, with no evidence of epilepsy found during prolonged EEG monitoring. In brief, the MEG seizure recordings were obtained in one patient with primary generalized absence epilepsy, in one patient with symptomatic generalized epilepsy, and in one patient with frontal lobe epilepsy. The iEEG seizure recordings were obtained from a patient with medically refractory temporal lobe epilepsy as part of the patient's routine clinical pre-surgical investigation. MEG recordings were obtained using a whole head CTF MEG system (Port Coquitlam, BC, Canada) with sensors covering the entire cerebral cortex, whereas iEEG electrodes were positioned in various locations including, in the temporal lobe epilepsy patient, the amygdala and hippocampal structures of both temporal lobes. EEG recordings were obtained using an XLTEK EEG system (Oakville, ON, Canada). The details of the acquisitions varied from patient to patient (e.g., acquisition rate varied from 200 to 625 Hz) and were taken into consideration for the data analyses. The duration of the recordings varied as well: for the seizure study, the MEG sample epochs were of 2 minutes duration each, with total recording times of 30-40 minutes; the iEEG patient sample was of 55 minutes duration. The sleep study data segments were each 2-4 minutes in duration, selected from continuous 24-hour recordings. Data analysis The only pre-processed data were those of the scalp EEG recordings. These were processed using a Laplacian derivation [11], to avoid the potential effects of the common reference electrode on synchro- nization [14] using the DSC algorithm [15]. Initially a phase synchrony index was calculated from all possible pairwise signal combinations, for which we use the standard procedure of estimating phase dif- ferences between two signals from the instantaneous phases extracted using the analytic signal concept via the Hilbert transform. To compute the synchrony index, several central frequencies, as specified in the text and figure legends, were chosen with a bandpass filter of 2 Hz on either side, hence, for one value of the central frequency f, the bandpass is f ± 2 Hz. The central frequencies were chosen according to the relevant behavioural states and some analytical limitations– thus for the sleep studies we choose 4 Hz (not lower because the extraction of the instantaneous phase was not optimal for central frequencies lower than 4). To see whether similar results were obtained with different frequencies, we chose others (see figures) provided there was power at those values. The phase synchrony index (R) was calculated using a 1-second running window and was obtained from the phase differences using the mean phase coherence statistic which is a measure of phase locking and is defined as R = hei∆θi where ∆θ is the phase difference between two signals. This analytical procedure has been described in great detail elsewhere [4, 5, 14]. The calculation of the index R was done for all possible signal pairs. The mean value of the R index thus obtained was then estimated for the desired time length. For instance, for the sleep recordings, the time period was the whole episode, which was, as noted above, between 2 and 4 minutes. For the seizure recordings, the periods to obtain the mean synchrony varied depending on the behavioural condition, for instance, in figure 1C the whole ictal event (labelled 'Sz') 3 and the initial 10-second portion of it ('10 sec Sz') were taken for the reason explained in the text and figure legend. The calculation of the number of connected signals and the entropy associated is described in the Results section. We note here that to assess entropy we assume that the different pairwise configurations are equiprobable, thus the entropy is reduced to the logarithm of the number of states, S = lnC (see notation in the main text). However, the estimation of C (the combinations of connections between diverse signals), is not feasible due to the large number of sensors; for example, for 35 sensors, the total possible number of pairwise connections is [1442] = 10296, then if we find in the experiment that, say, 2000 pairs are connected, the computation of [102962000] has too large numbers for numerical manipulations, as they cannot be represented as conventional floating point values in, for instance, MATLAB. To overcome this difficulty, we used the well-known Stirling approximation for large n : ln(n!) = n ln(n)n. The Stirling approximation is frequently used in statistical mechanics to simplify entropy-related computations. Using this approximation, and after some basic algebra, the equation for entropy reads, S = N ln(N/N − p) − p ln(p/N − p), where N is the total number of pos- sible pairs of channels and p the number of connected pairs of signals in each experiment (see Results for details and notation). Because this equation is derived from the Shannon entropy, it indicates the information content of the system as well (37). In addition to entropy, we used another measure of complexity, the Lempel-Ziv (L-Z) complexity, based in the Kolmogorov deterministic complexity [16]. This complexity measure the amount of non- redundant information in a string by estimating the minimal size of the "vocabulary" necessary to describe the string [17]. Strings with high L-Z complexity require a large number of different patterns ("words") to be reproduced, while strings with low complexity can be largely compressed with a few patterns employed to eliminate redundancy with no loss of information. For this purpose, values of the matrix B (defined in Results) were placed in a one-dimensional vector and its L-Z complexity determined. 3 Result Guided by proposals that consciousness requires medium values of certain features of cell assemblies, e.g. not too high or low synchrony or correlations [18, 19, 3], or halfway between order and disorder [20], we chose to quantify the number of possible configurations the brain can adopt in different behavioural conditions. Our basic approach consists in the estimation of the number of possible pairwise connections between recorded brain signals. Signals included MEG, iEEG and scalp EEG recordings; details can be found in Methods. We are limited to pairwise combinations of the signals because of the manner in which phase synchrony is computed – as phase differences between two signals – and we use phase synchronization as the means to determine "connectivity" between the two signals. Once the number of "connected" signals is known, we estimate the entropy of those pairwise combinations. The results obtained with recordings acquired during conscious states are compared with those acquired during unconscious states, which included sleep (all stages) and epileptic seizures. To determine connectivity, we use an accepted approach of computing a phase synchronization index (details in Methods). It must be noted that, while many studies use the words synchrony and connectivity as synonymous, in reality phase synchrony analysis reveals only a correlation between the phases of the oscillations between two signals, and not a real connectivity which depends on several other factors (this matter has been covered in detail in [21, 22]. Nevertheless, due to the unfeasibility of an accurate, realistic 4 estimation of connectivity which would necessitate individual cell recordings from entire cell ensembles as well as structural connectivity details, we use an accepted version under the assumption that the phase relations may represent, at least, some aspect of a functional connectivity. Hence, in order to evaluate interactions ("connections"), we take each sensor/channel as one "unit", and define a pair of signals as "connected" if the phase synchrony index is larger than a threshold. The threshold is determined for each individual, and is the average synchrony index in the most normal alert state, the 'awake eyes-open' condition, when the individual is fully alert and processing the sensorium in a regular fashion. Because our data include the three recording methodologies aforementioned, we have the opportunity to assess the reproducibility of the results in various types of recordings. While we work at the signal level we will make the reasonable assumption that the MEG and scalp EEG sensors record cortical activity underlying those sensors and thus throughout the text we will use the terms brain signals or brain areas/networks as synonymous. The iEEG, obviously, records signals at the source level. Note that we are not interested in the specific pattern of connectivity among brain sources/areas, but rather in the global states. These points are further discussed in the Discussion section. Phase synchronization for specific frequencies (details in Methods) is calculated for each pair of channels and a "connectivity" matrix S is obtained, whose entries are the average values of the syn- chrony index during a certain time period for each pairwise configuration. From this matrix, a Boolean connectivity matrix B is calculated, with 0 entry if the corresponding synchrony index is lower than a threshold, and 1 if higher. We define two channels as "connected" if the corresponding entry in matrix B is 1. Then we use the combinations of connected channels as a 'complexity' measure. The total number of possible pairs of channels given a specific channel montage is given by the binomial coefficient N = N c!/2! (N c − 2)! where N c is the total number of channels in the recording montage, normally 144− 146 in the case of MEG sensors, and between 19 and 35 with iEEG and scalp EEG. For instance, in our MEG recordings we haveN c = 144, thus N = 10, 296possible pairs of connected sensors are obtained. For each subject we calculate p, the number of connected pairs of signals in the different behavioural states, using the aforementioned threshold of the synchrony index (which varies for each subject), and estimate C, the number of possible combinations of those p pairs, using the binomial coefficient again: C = N!/p!(N − p)! where N is the aforesaid value. In sum, all these calculations represent the relatively simple combinatorial problem we are trying to solve: given a maximum total of N pairs of connected signals, in how many ways can our experimental observation of p connected pairs (that is, the number of 1's in matrix B) be arranged. We then compute the entropy and Lempel-Ziv complexity associated with those p values. Figure 1 depicts the entropy (S), which, assuming equiprobable states, is the logarithm of the number of states, S = ln C (see Methods for the estimation of entropy using very large C values) in three epileptic patients. We note that this equation using the natural logarithm allows for a calculation of both the Gibbs and the Shannon entropy, if needed, which differ from S by a constant multiplicative factor k (the Boltzmann constant) and 1/(ln2) respectively. In reality the entropy estimation does not provide any further information, as the main, crucial result is the number of configurations, C. However, we have done it since it is a standard manner to quantify the "complexity" of the number of microstates. The entropy data points are graphed on the curve that represents the entropy of all points in the binomial distribution, where the maximum number of configurations (that is, maximum entropy) occurs in the middle of the graph. Note in that figure that during conscious states, when patients are not having generalised seizures with loss of awareness, the entropy is close to the maximum, whereas entropy is lower (more distant from the top) for the seizure states. The values during the seizures fall 5 on the right-hand side of the graph because, due to the higher synchrony during ictal (seizure) events such that the number of coupled channels (note that the x-axis is the number of connected signals) is larger than during interictal (between seizure) activity, there are fewer pairwise configurations and thus lower entropy. This phenomenon seems associated with the level of consciousness since when the seizures are not generalised (Figure 1C and 1D), and the patients remain responsive and conscious, the entropy values are similar to those of interictal (baseline) activity. Where in the curve the data points are located depends of course on the synchrony index. Because seizures had higher synchrony than interictal periods ("baseline"), the number of coupled signals is greater and the number of combinations is lower. We observe similar trends in the case of sleep. Figure 2 depicts some examples. Note how during wakefulness the entropy is closer to the maximum of the curve, whereas the deeper the sleep stage, the more distant from the maximum the values are. The deepest sleep stage, slow wave 3-4 ('sws3-4'), has consistently the lowest entropy. Interestingly, the entropy during REM sleep is very close, in most cases, to the normal, alert state. This result may not be as surprising as it sounds if we consider the mental activity during REM episodes that are normally associated with dreams. It is worth noting too that in recordings taken when the subjects had their eyes closed, the entropy is much lower than during the eyes open condition, and sometimes it is as low as during slow-wave 3-4 sleep. The results for two central frequencies (4 and 8 Hz) are shown in Figure 2A and 2C, to demonstrate that it is not always the case of high synchrony having lower entropy; sometimes it is lower synchrony (e.g., results at 8 Hz) that resulted in fewer channel combinations and thus lower entropy as compared to fully alert states. In Figures 1 and 2 we have shown results with iEEG and MEG (Figure 1) or iEEG and scalp EEG (Figure 2) recordings to demonstrate that the same qualitative results are obtained with these three recordings techniques. Thus these results do not depend on recording methodology. Shown in Figure 3 is an example of the time course of the entropy before and during an ictal event. Hence, we have demonstrated that the specific values of the synchrony index R in fully alert states represent the largest number of combinations (microstates, see Discussion) of pairwise signal configurations. This method solves the potential problem of the interpretation of the different R values obtained with various recordings techniques. For example, in our experiments, average values in baseline conditions are 0.248 ± 0.2 for MEG, 0.428 ± 0.04for iEEG, and 0.46 ± 0.05 for scalp EEG, nevertheless, in our study the number of combinations are "normalised" to the number of recording sensors and show a final result (the entropy) that is independent of the structure and synchrony magnitudes of the recording methodology. The main idea derived from these results is represented in Fig. 4. To further explore whether the decrease in entropy has a parallel with a decrease in other forms of complexity, the Lempel-Ziv complexity of the number of configurations was assessed. Tables 1 and 2 illustrate the Lempel-Ziv complexity estimated for the B matrices, where a complementary result can be seen: unconscious states have lower values of complexity. A decrease of complexity in the raw electrophysiological signals was obtained too (data not shown), hence this may be a phenomenon observable at different levels of description. 4 Discussion Our attempts at seeking features of brain organization that allow for adequate processing of sensory stimuli have provided evidence that a greater number of possible configurations of interactions between 6 brain networks is associated with alert states, representing high entropy associated with the number of those combinations, whereas lower entropy (and thus fewer combinations of connections) is characteristic of either unconscious states or altered states of alertness (eyes closed). This observation reflects a relatively simple general organising principle at this collective level of description, which results in the emergence of properties associated with consciousness. With the advent of the 'Big Data' era and the related torrent of empirical observations, the search for organising principles that result in the emergence of biological phenomena seems more crucial than ever. We tried to uncover relatively simple laws that capture the bounds in the global organization of a biological system that enable it to become adaptable (i.e., responsive) to an environment, or, in neuroscientific terms, the features of optimal brain organization (in terms of connections) that allow brains to adequately process sensory stimuli. We focused on the global states and did not investigate specific patterns of connectivity between brain areas as a variety of other studies have assessed these inter-regional interactions in conscious and unconscious states [23, 24]. The fact that our results are similar, independent of recording methodology, demonstrates the robustness of the analysis. On that note, we remark that while we have used the term 'connectivity', in reality the analysis reveals only correlation between phases of oscillation, as already discussed in Results. The present study can be considered an extension of previous work where it was proposed that a general organising principle of natural phenomena is the tendency toward a maximal, or more probable, distribution of energy [7], which can be encapsulated by the notion of the maximization of information transfer [25]. As well, the notions of information and energy exchange are conceptually related: "the common currency paying for all biological information is energy flow" [26, 9]. In the final analysis, information exchange implies energy exchange, hence we interpret information exchange as energy re- distribution as proposed in [7], even though our study is focused not on energy considerations but on the number of states. Other studies have assessed the importance of brain synchronization to optimally transfer information [27]. We interpret our observation that the number of pairwise channel combina- tions – that we take as interactions/connections between brain networks – occurs near the maximum of possible configurations in periods with normal alertness, as that greater number of configurations of interactions represents the most probable distribution of energy/information resulting in conscious awareness. The configuration entropy we calculate measures the information content of the functional network, and has been used in other works for the purpose of quantifying information [28]. One some- what surprising result is the low entropy during the eyes-close condition. One could argue that having the eyes closed does not change much the conscious state, and yet we observe reduced entropy as com- pared with eyes-open condition. Hence our observations may indicate not only states of consciousness but also optimality of sensory processing – considering the great importance of visual processing in humans, interrupting visual inputs should result in considerable changes in the dynamics; for instance, an apparent alteration of brain dynamics upon eye closure is the appearance of alpha waves most clearly in parieto-occipital regions. Interesting too is the relatively high entropy associated with REM episodes, perhaps a reflection of the partial awareness during dreams. Perhaps the main difference between dream and awake states is psychological, but they share similar brain dynamics. It has been proposed that aspects of awareness emerge when certain levels of complexity are reached [29]. It is then possible that the organization (complexity) needed for consciousness to arise requires the maximum number of configurations that allow for a greater variety of interactions between cell assemblies because this structure leads to optimal segregation and integration of information, two fun- damental aspects of brain information processing [20] . Our results help cast the study of consciousness 7 and cognition into more of a physics framework that may provide insight into simple principles guiding the emergence of conscious awareness, and perhaps the proposed thermodynamics for a network of connected neurons [18] can be extended to explain cognition. Some classical studies [8] have already characterized biological order in terms of functions of the number of states of a system. It is tempt- ing to link our observations with the typical chemical equilibrium that, despite being composed of a myriad of microstates, when viewed at the macroscopic level produces some useful laws, like the law of mass action describing chemical balance, and thus the perspective we develop here may help guide research to uncover organising principles in the neurosciences. In physics, microstates that yield the same macrostate form an ensemble. A system tends to approach the most probable state, maximising entropy under present constraints, and the resulting macrostate will be represented by the maximum number of microstates. Hence, the macrostate with higher entropy (see scheme in Fig. 4) we have defined, composed of many microstates (the possible combinations of connections between diverse net- works, our C variable defined in Results), can be thought of as an ensemble characterised by the largest number of configurations. Here we define an ensemble of microstates as all possible configurations of connectivity leading to the same macrostate (having the same number of connected pair of signals, p). The entropy of this macrostate, given by the logarithm of the number of combinations (our C), is the number of microstates that are compatible with the given macrostate (assuming all microstates have same statistical weight). In neurophysiological terms, each microstate represents a different connectiv- ity pattern and thus is associated with, in principle, different behaviours or cognitive processes. The macrostate that we find associated with wakeful normal states (eyes open) is the most probable because it has the largest entropy (largest number of combinations of connections). While many elementary cel- lular microscopic processes are far from equilibrium (e.g., ionic gradients), at the macroscopic level the system tends towards equilibrium, as most natural phenomena remain in near-equilibrium conditions [30]. At the same time, the ensemble of microstates associated with normal sensory processing features the most varied configurations and therefore offers the variability needed to optimally process sensory inputs. In this sense, our results support current views on the metastability of brain states [1] in that the states should not be too stable for efficient information processing, hence the larger the number of possible interactions, the more variability is possible. Equally, the results are consistent with the global workspace theory [31] in that the most widespread distribution of information leads to conscious awareness. Furthermore, these observations relate as well to the information integrated theory [32], in that consciousness increases in proportion to the system's repertoire of states, thus the more combinations possible, the more states available, and here we can de- fine states as configurations of interactions. Additionally, the results support computational/theoretical studies showing that patterns of organised activity arise from the maximization of fluctuations in syn- chrony [33], and that transitions between conscious states are achieved by just varying the probability of connections in neural nets [34]. In general, our observations highlight the fundamental importance of fluctuations in neuronal activity as the source of healthy brain dynamics. More specifically, our results have a precise parallel with the work of Hudetz et al. (2014), where they graph a dispersion index versus an activation level (their figure 7B) and propose that consciousness resides at the top of the curve, and anaesthetic states and seizures to lower and higher activation levels respectively, as we show in Figures 1 and 2. Their 'activation level' could correspond to the number of signals that take part in the combinations (our x axes in the graphs), and their 'dispersion index' to our number of combinations (the y axes). Other studies have proposed as well that consciousness requires medium values of certain features of cell assemblies [32, 35]. 8 In sum, along with others, we consider cognition/consciousness not a static property but a dynamic process with constant flux of energy, or information exchange [25]. Even though we have talked above about a macrostate, this should not be taken as a fixed state, rather it contains dynamic processes represented by the microstates, the re-arrangements of connections among brain cell ensembles. The emergent features of cognitive phenomena that can be termed "conscious" arise once an efficient web of connections endowed with certain complexity appears. The fact that values of phase synchrony during fully alert states gave us the largest entropy of the number of pairwise signal combinations explains in part the neurophysiological organization underlying these specific values of the quantification of brain synchrony. Studies at this level of description may help to understand how consciousness arises from organization of matter. In our view, consciousness can be considered as an emergent property of the organization of the (embodied) nervous system, especially a consequence of the most probable distribution that maximizes information content of brain functional networks. On a technical note, we remark that in our analysis we had to choose a threshold to define "connec- tivity", and therefore a suitable baseline had to be chosen. This procedure is obviously biased in that a signal corresponding to a baseline has to be ascribed as the one providing the threshold, and we chose the signals corresponding to the (psychological) state that is most suited for the purposes of adaptabil- ity: fully alert and receiving all sensory inputs (awake with open eyes). By this choice, and according to our methods, that signal is already ascribed as having maximum entropy. We tried another less prejudiced method, using surrogates of the original signals, and then computing the average synchrony index among the surrogate population (10 phase-randomised surrogates per original channel/signal) that is the threshold to define connectivity. It turns out that the value of the magnitude of synchrony of the surrogates is close to the one for the aforementioned baseline chosen, so the results do not vary; nev- ertheless this new method still assigns the largest entropy to the random signals (surrogates), so there is still the assumption that the average synchrony of the stochastic signals is a good approximation to define connections among brain networks. Our current purpose is to completely avoid using thresholds of synchrony indices, for which we are presently working on a scheme that assigns connectivity in the time domain. 5 Conclusions and conjectures on the structure of brain-behaviour- environment It is tempting to speculate, based on these results and the conclusions of a previous study [7] that there could be a universal logic ruling the evolution of natural phenomena - biological and nonbiological- and the nervous system in particular: patterns emerge from a central theme captured by maximising information exchange. Because, in the final analysis, all exchange of information implies exchange of energy, natural phenomena tend towards the most probable distribution of energy, and thus the interactions among system constituents tends to be maximized. Because the brain functions to maintain a predictive model of the environment (the reason the brain evolved is to model the environment, after all), then perhaps the brain's global configuration has to "copy" what is out there: and out there energy distributes in all possible microstates (second principle of thermodynamics). Then to process such variability in nature, the nervous system should have same structure, and the result is the 'inverted U' that has appeared in our analysis and has been theoretically proposed in other publications [20], the top of the curve representing more possible 9 combinations to handle information/energy exchanges. On the other hand, in the extremes of this curve we find fewer microstates, thus these are not optimal situations to process the many microstates in the environment. The key then is not to reach the maximum number of units interacting (which would be all-to-all connections and thus only one possible microstate), but rather the largest possible number of configurations allowed by the constraints. In a similar fashion, it has been argued that the brain needs to show criticality because natural phenomena possess critical dynamics [36]. Then, perchance, consciousness can be considered as an emergent property of the organization of the embodied nervous system submerged in an environment, consequence of the most probable distribution of energy (information exchange) in the brain. In this regard, consciousness (like biochemistry) may represent thus an optimal channel for accessing sources of (free) energy. 10 References [1] JA. S. Kelso. Dynamic patterns: The self-organization of brain and behavior. MIT press, 1997. [2] D. D Garrett, G. R. Samanez-Larkin, S. WS. MacDonald, U. Lindenberger, A. R McIntosh, and C. L Grady. Moment-to-moment brain signal variability: a next frontier in human brain mapping? Neuroscience & Biobehavioral Reviews, 37(4):610–624, 2013. [3] V. Nenadovic, J. L. Perez Velazquez, and J. S. Hutchison. Phase synchronization in electroen- cephalographic recordings prognosticates outcome in paediatric coma. PloS one, 9(4):e94942, 2014. [4] F. Mormann, K. Lehnertz, P. David, and C. E. Elger. Mean phase coherence as a measure for phase synchronization and its application to the eeg of epilepsy patients. Physica D: Nonlinear Phenomena, 144(3):358–369, 2000. [5] L.. Garcia Dominguez, R. A. Wennberg, W. Gaetz, D. Cheyne, O. C. Snead, and J. L. Perez Ve- lazquez. Enhanced synchrony in epileptiform activity? local versus distant phase synchronization in generalized seizures. The Journal of neuroscience, 25(35):8077–8084, 2005. [6] D. C. Shields, J. W. Leiphart, D. L. McArthur, P. M. Vespa, M. L. Quyen, J. Martinerie, and J. R. Soss. Cortical synchrony changes detected by scalp electrode electroencephalograph as traumatic brain injury patients emerge from coma. Surgical neurology, 67(4):354–359, 2007. [7] J. L. Perez Velazquez. Finding simplicity in complexity: general principles of biological and non- biological organization. Journal of biological physics, 35(3):209–221, 2009. [8] H. J. Morowitz. Some order-disorder considerations in living systems. The bulletin of mathematical biophysics, 17(2):81–86, 1955. [9] E. Smith. Thermodynamics of natural selection i: Energy flow and the limits on organization. Journal of theoretical biology, 252(2):185–197, 2008. [10] J.J. Wright and D.T.J. Liley. Dynamics of the brain at global and microscopic scales: Neural networks and the eeg. Behavioral and Brain Sciences, 19(02):285–295, 1996. [11] P. L. Nunez. Toward a quantitative description of large-scale neocortical dynamic function and eeg. Behavioral and Brain Sciences, 23(03):371–398, 2000. [12] J. L. Perez Velazquez, Garcia Dominguez L., V. Nenadovic, and R. A. Wennberg. Experimen- tal observation of increased fluctuations in an order parameter before epochs of extended brain synchronization. Journal of biological physics, 37(1):141–152, 2011. [13] R. Wennberg. Intracranial cortical localization of the human k-complex. Clinical Neurophysiology, 121(8):1176–1186, 2010. [14] R. Guevara, J. L. Pérez Velazquez, V. Nenadovic, R. Wennberg, G Senjanović, and L. G. Dominguez. Phase synchronization measurements using electroencephalographic recordings. Neu- roinformatics, 3(4):301–313, 2005. 11 [15] J. Kayser and C. E. Tenke. Principal components analysis of laplacian waveforms as a generic method for identifying erp generator patterns: I. evaluation with auditory oddball tasks. Clinical neurophysiology, 117(2):348–368, 2006. [16] A. N. Kolmogorov. Three approaches to the quantitative definition of information. Problems of information transmission, 1(1):1–7, 1965. [17] A. Lempel and J. Ziv. On the complexity of finite sequences. Information Theory, IEEE Transac- tions on, 22(1):75–81, 1976. [18] G. Tkacik, T. Mora, O. Marre, D. Amodei, II Berry, J. Michael, and W. Bialek. Thermodynamics for a network of neurons: Signatures of criticality. arXiv preprint arXiv:1407.5946, 2014. [19] V Nenadovic, J. S. Hutchison, L. G. Dominguez, H. Otsubo, M. P. Gray, R Sharma, J. Belkas, and J. L. Perez Velazquez. Fluctuations in cortical synchronization in pediatric traumatic brain injury. Journal of neurotrauma, 25(6):615–627, 2008. [20] G. Tononi, G. M. Edelman, and O. Sporns. Complexity and coherency: integrating information in the brain. Trends in cognitive sciences, 2(12):474–484, 1998. [21] J. L. Perez Velazquez, R. G. Erra, R. Wennberg, and L. G. Dominguez. Correlations of cellular activities in the nervous system: physiological and methodological considerations. In Coordinated Activity in the Brain, pages 1–24. Springer, 2009. [22] J. L. Perez Velazquez and M. Frantseva. The Brain-Behavior Continuum: The Subtle Transition Between Sanity and Insanity. World Scientific, 2011. [23] G Dumermuth and D Lehmann. Eeg power and coherence during non-rem and rem phases in humans in all-night sleep analyses. European neurology, 20(6):429–434, 1981. [24] K. Wang, M. L. Steyn-Ross, D.A. Steyn-Ross, M. T. Wilson, and J. W. Sleigh. Eeg slow-wave coherence changes in propofol-induced general anesthesia: experiment and theory. Frontiers in systems neuroscience, 8, 2014. [25] Hermann Haken. Information and self-organization: A macroscopic approach to complex systems. Springer Science & Business Media, 2006. [26] H. J Morowitz. Energy flow in biology; biological organization as a problem in thermal physics. 1968. [27] A Buehlmann and G Deco. Optimal information transfer in the cortex through synchronization. PLoS Comput Biol, 6(9):e1000934, 2010. [28] W. Bialek, F. Rieke, R de Ruyter van Steveninck, and D Warland. Spikes: exploring the neural code. MIT. Roddey, JC, Girish, B., & Miller, JP (2000). Assessing the performance of neural encoding models in the presence of noise. Journal of Computational Neuroscience, 8(95):112, 1997. [29] M. Gell-Mann. Consciousness, reduction, and emergence. Annals of the New York Academy of Sciences, 929(1):41–49, 2001. 12 [30] I. Prigogine. Introduction to thermodynamics of irreversible processes. New York: Interscience, 1967, 3rd ed., 1, 1967. [31] B. J. Baars. A cognitive theory of consciousness. Cambridge University Press, 1993. [32] G. Tononi. An information integration theory of consciousness. BMC neuroscience, 5(1):42, 2004. [33] V. Vuksanović and P. Hövel. Dynamic changes in network synchrony reveal resting-state functional networks. Chaos: An Interdisciplinary Journal of Nonlinear Science, 25(2):023116, 2015. [34] D. W. Zhou, D. D. Mowrey, P. Tang, and Y. Xu. Percolation model of sensory transmission and loss of consciousness under general anesthesia. Physical review letters, 115(10):108103, 2015. [35] A. A Fingelkurts, A. A. Fingelkurts, S. Bagnato, C. Boccagni, and G. Galardi. Do we need a theory-based assessment of consciousness in the field of disorders of consciousness? Frontiers in human neuroscience, 8:402, 2014. [36] D. R. Chialvo. Emergent complex neural dynamics. Nature physics, 6(10):744–750, 2010. 13 State Subject #1 Baseline Seizure Subject #2 Baseline Seizure Subject #3 Baseline Seizure Subject #4 Baseline Seizure L-Z complexity 0.7 0.2 0.61 0.39 0.86 0.71 0.61 0.61 Table 1: Values of Lempel-Ziv (L-Z) complexity derived from the string of connections (details in Meth- ods) in conscious (baseline) and unconscious (seizure) states, in four patients. Note lower complexity during seizures in patients # 1 − 3; patient # 4 (Figure 1D) did not have fully generalised seizures. Subject #1 Alert eyes open State SWS 2 SWS 3-4 REM Subject #2 Alert eyes open Alert eyes close SWS 3-4 Subject #3 Alert eyes open Subject #4 Alert eyes open Alert eyes closed SWS 2 SWS 3 REM Subject #5 Alert eyes open iEEG L-Z complexity 0.81 0.42 0.018 0.73 0.94 0.94 0.33 N/A N/A N/A N/A 0.88 0.55 0.57 0.05 0.6 N/A N/A N/A N/A N/A scalp L-Z complexity 1.01 1.0 0.0 0.96 1.06 0.85 0.22 1.07 1.07 1.07 0.97 1.0 1.0 9.8 0.0 1.06 0.97 1.08 1.04 0.8 1.12 SWS 2 SWS 3-4 REM SWS 1 SWS 2 SWS 3-4 REM Table 2: Values of Lempel-Ziv (L-Z) complexity derived from the string of connections in different sleep stages. When both recordings were obtained from a subject, both iEEG and scalp EEG data were analysed (subjects # 3 and # 5 did not have iEEG recordings). The L-Z complexity is consistently lower, regardless of recording methodology, in the deepest sleep stage (SWS 3-4). These results parallel those of entropy estimations shown in figure 2 14 Figure 1: Graphs representing the entropy of the number of pairwise configurations of signals in epileptic patients during conscious (baseline) and unconscious (generalised seizure) states. A, derived from MEG recordings in a patient with primary generalised epilepsy, shows entropy associated with a normal alert period (baseline, 'Base') and a generalised absence seizure ('Sz'), estimated from synchrony values at two central frequencies (defined in Methods) of 5 and 12 Hz. The curve in this and other graphs here and in Figure 2 represents the possible entropy values of all possible numbers of pairwise combinations, yielding an inverted U or, for very large numbers, a Gaussian. Note that here as well as in all generalised seizures analysed, the entropy values associated with alert, baseline conditions were closer to the maximum (top of the curve) than those associated with the seizures. B, entropy values of two seizures and their corresponding baseline ('Base') activity (computed using a time period of 30-40 minutes before the ictus) in a patient with secondary (symptomatic) generalised epilepsy (MEG recordings). C, derived from iEEG recordings in a patient with temporal lobe epilepsy, shows the entropy during the alert state ('baseline'), during the initial 10 seconds of the seizure when the patient was still responsive and alert ('10 sec Sz'), and during the rest of the seizure when it became generalised and the patient was unresponsive ('Sz'). Note that when the ictus has not yet generalised, the entropy is similar to that of normal alertness. D, another example of a non-generalised seizure in a patient with frontal lobe epilepsy (MEG recordings). 15 2000400060008000100001000300050007000BaselineSz2000400060008000100001000300050007000S (entropy)Bae zs 5HSz @ 5 HzBase@ 12 HzSz @ 12 HzABase #2Sz # 12000400060008000100001000300050007000Base #1Sz # 2B10 sec Sz5101520253035510152025BaselineSzCDS (entropy)Number of coupled channels Same graph types as in Figure 1, using sleep recordings. Figure 2: In each subject, data samples were of 2-4 minutes duration during wakefulness with eyes open ('awake') or eyes closed ('Eyecl'), and sleep stages slow-wave 2 ('Sws2'), slow-wave 3-4 ('Sws3-4') and rapid eye movement ('REM'). A, results derived from iEEG recordings in a subject investigated with bilateral frontal and temporal electrodes and simultaneous scalp EEG. Entropy estimated from synchrony values at two central frequencies of 4 and 8 Hz. As occurred in the patient recordings shown in Figure 1, the baseline, alert state (in this case labelled 'awake') is closer to the top of the curve, having greater entropy. The deepest sleep stage, slow wave 3-4 ('sws3-4'), has the lowest entropy. B, same subject but using the scalp EEG recordings for the calculations, showing similar trend (evaluated at central frequency of 4 Hz). C, results derived from a different subject investigated with right frontal electrodes and simultaneous scalp EEG, with synchrony evaluated at two central frequencies of 4 and 8 Hz using the iEEG signals. Note that the eyes closed ('Eyecl') condition has lower entropy than that of the normal alert state with open eyes. Depending on the frequency of analysis, the entropy in 'Eyecl' falls toward the left or right side of the curve; e.g., at 8 Hz, because the synchrony is higher (more coupled channels) due to alpha waves at 8-10 Hz and the entropy is reduced due to fewer combinations of connections, as occurs similarly during seizures (Figure 1). D, same subject but using the scalp EEG signals, showing similar results to those obtained with iEEG (evaluated at central frequency of 4 Hz). 16 200400600800100012000100200300400500600700800S (entropy)awakeSws2Sws3-4REMawakeSws2Sws3-4REM4 Hz8 Hz*AawakeSws2Sws3-45010015020025030050100150200250REMBREM10020030040050060050100150200250300350400awakeEyeclSws2Sws3-4awakeEyeclREMSws2Sws3-44 Hz8 Hz*CREM5010015020025030050100150200250awakeEyeclSws2Sws3-4DS (entropy)Number of coupled channels Figure 3: Time course of the entropy of the number of configurations of connected MEG signals before, during and after a generalised absence seizure. MEG signal from one channel is shown at top, the ictus ('Sz') occurs towards the end of the 2 minute recording. Notice the drop in entropy during the seizure. 17 SzTime (sec)600040002000329664168048112S (entropy) Figure 4: Proposed general scheme of the relation between global brain connectivity and behavioural states. Normal alertness resides at the top of the curve representing the number of configurations of connections the system can adopt, or the associated entropy. The maximisation of the configurations (microstates) provides the variability in brain activity needed for normal sensorimotor action. Abnor- mal, or unconscious states, are located farther from the top, and are characterised by either large or small number of "connected" networks therefore exhibiting lower number of microstates (hence lower entropy) that are not optimal for sensorimotor processing. 18 number networks "connected"conscious/alertunconsciousEntropy (S) or number of configurations low Shigh SÞnumber microstates
1804.01906
3
1804
2018-05-23T12:40:58
An Accelerated LIF Neuronal Network Array for a Large Scale Mixed-Signal Neuromorphic Architecture
[ "q-bio.NC", "cs.ET", "physics.bio-ph", "physics.comp-ph" ]
We present an array of leaky integrate-and-fire (LIF) neuron circuits designed for the second-generation BrainScaleS mixed-signal 65-nm CMOS neuromorphic hardware. The neuronal array is embedded in the analog network core of a scaled-down prototype HICANN-DLS chip. Designed as continuous-time circuits, the neurons are highly tunable and reconfigurable elements with accelerated dynamics. Each neuron integrates input current from a multitude of incoming synapses and evokes a digital spike event output. The circuit offers a wide tuning range for synaptic and membrane time constants, as well as for refractory periods to cover a number of computational models. We elucidate our design methodology, underlying circuit design, calibration and measurement results from individual sub-circuits across multiple dies. The circuit dynamics match with the behavior of the LIF mathematical model. We further demonstrate a winner-take-all network on the prototype chip as a typical element of cortical processing.
q-bio.NC
q-bio
1 An Accelerated LIF Neuronal Network Array for a Large Scale Mixed-Signal Neuromorphic Architecture Syed Ahmed Aamir∗, Student Member, IEEE, Yannik Stradmann∗, Paul Muller, Christian Pehle, Andreas Hartel, Andreas Grubl, Johannes Schemmel, Member, IEEE, and Karlheinz Meier, Member, IEEE Abstract-We present an array of leaky integrate-and-fire (LIF) neuron circuits designed for the second-generation BrainScaleS mixed-signal 65-nm CMOS neuromorphic hardware. The neu- ronal array is embedded in the analog network core of a scaled- down prototype HICANN-DLS chip. Designed as continuous- time circuits, the neurons are highly tunable and reconfigurable elements with accelerated dynamics. Each neuron integrates input current from a multitude of incoming synapses and evokes a digital spike event output. The circuit offers a wide tuning range for synaptic and membrane time constants, as well as for refractory periods to cover a number of computational models. We elucidate our design methodology, underlying circuit design, calibration and measurement results from individual sub- circuits across multiple dies. The circuit dynamics match with the behavior of the LIF mathematical model. We further demonstrate a winner-take-all network on the prototype chip as a typical element of cortical processing. Keywords-Analog integrated circuits, Neuromorphic, Leaky Integrate and Fire, 65nm CMOS, Spiking neuron, OTA, Opamp, Tunable resistor, Winner-take-all network I . I N T R O D U C T I O N T HE architecture of digital microprocessors is fundamen- tally different from that of the central nervous system. While the brain is a massively parallel structure of neurons in- terconnected through synapses [1], microprocessors are mostly based on a von Neumann architecture [2], [3] with logic gates as the elementary primitives. The human brain consumes only approximately 20 W [4], while its performance as a general- purpose problem solver is still unmatched by any computer algorithm. Taking inspiration from this biological feat, neuromorphic architectures not only adopt a non-von Neumann architecture by collocating memory close to the computational element, but also introduce massive parallelism, high energy efficiency, reconfigurability, fault tolerance, and integrate computational *Both authors contributed equally to this work. Manuscript submitted August 29, 2017, revised January 20, 2018; April 4, 2018; May 14, 2018, accepted May 21, 2018. This work has received funding from the European Union Seventh Framework Programme ([FP7/2007-2013]) under grant agreement nos. 604102 (HBP), 269921 (BrainScaleS), the Horizon 2020 Framework Programme ([H2020/2014-2020]) under grant agreement no. 720270 (HBP) as well as from the Manfred Stark Foundation. All authors are with the Kirchhoff Institute for Physics, Heidelberg Univer- sity, D-69120 Heidelberg, Germany. Email: {aamir,yannik.stradmann}@kip.uni-heidelberg.de. models of neural elements using CMOS technologies [5]– [7]. They compute at biological timescales or at a specified speed-up ("acceleration") factor. In particular, analog imple- mentations integrate bio-physically inspired neuron models as computational elements in order to capture the rich temporal dynamics of the neuronal membrane. Since their emergence in the late eighties [5], neuromorphic circuits and systems have evolved with a variety of large-scale architectures, model implementations and technologies [4], [8]. For example, [9]–[14] describe systems that adhere to the ini- tial idea of exploiting weak-inversion characteristics of CMOS circuits to emulate ion flow dynamics. More recently, with the advancements in cognitive computing driven in part by the speech and object recognition made possible by deep neural networks, major industrial players [15], [16] explore neuro- morphic architectures and integrate digital phenomenological neuron models in ultra deep-submicron technologies - the most recent being Intel's 14 nm FinFET Loihi chip [16]. Within EU's framework for brain-inspired computing [17], [18], the SpiNNaker and BrainScaleS systems present large- scale neuromorphic platforms. While SpiNNaker [19] uses arrays of many-core interconnected ARM-based microproces- sor nodes and implements a software-defined neuron model, our approach as described within the BrainScaleS hardware system [20], [21] is a wafer-scale implementation with an analog physical neuron model and accelerated-time operation. The system utilizes an entire post-processed CMOS wafer for large scale integration by interconnecting multiple identical on- wafer chips [21]. In the second-generation BrainScaleS architecture [22], [23], we present, for the first time, a mixed-signal neuromorphic computing core that integrates a custom Single Instruction Multiple Data (SIMD) processor with the Analog Network Core (ANC) (comprised of the neuron array and synapse matrix) in 65 nm CMOS. We have previously described the plasticity subsystem of the chip [23], [24]. Here we present the first silicon implementation and characterization of the neuronal array [25] based on the Leaky Integrate and Fire (LIF) neuron model. We demonstrate its high tunability, reconfig- urability, reliable parameter mapping through calibration and the network operation on a scaled-down prototype chip. The neuron array is tested in tight integration with the synapses and dedicated analog bias storage [26], [27] providing voltage and current biases for tuning the neuron's behavior. The presented chip features a decreased neuronal and synaptic count (32 ©2018 IEEE. Personal use of this material is permitted. Permission from IEEE must be obtained for all other uses, in any current or future media, including reprinting/republishing this material for advertising or promotional purposes, creating new collective works, for resale or redistribution to servers or lists, or reuse of any copyrighted component of this work in other works. neurons interconnected with 32×32 synapses instead of 512 neurons and 256×512 synapses), but integrates all essential elements of the scaled-up version (see Sec. II). We demonstrate a winner-take-all network as an example application. The detailed circuit description of synapses, the analog bias storage and the plasticity mechanism of the chip is omitted in this work. In the subsequent sections, we describe the chip architecture of the HICANN-DLS prototype, as well as the design pre- considerations (Sec. II & Sec. III). The design, measurement, calibration and performance results of the neuron as well as of the individual sub-circuits are described in Sec. IV. A demonstration of a winner-take-all network is detailed in Sec. V. We discuss and compare our results with other large- scale neuromorphic systems in Sec. VI before concluding the paper in Sec. VII. I I . T H E H I C A N N - D L S C H I P The BrainScaleS wafer-scale system [21], [28] is built in the first phase with 180 nm CMOS HICANN neuromorphic chips, operated 104–105 times faster than biological timescale. The second-generation hardware features the enhanced High Input Count Analog Neural Network with Digital Learning System (HICANN-DLS) chips as the fundamental building block [22]. HICANN-DLS is a 65 nm CMOS mixed-signal system on-chip solution with integrated analog and digital cores and operated one thousand times faster than biological real-time. The chip is under development and an enhanced scaled-up version of the current prototype will replace the HICANN chip in the second-generation wafer-scale platform. A simplified architecture of the initial chip prototype con- nected to its off-chip measurement system is sketched in Fig. 1a. The left half shows the ANC arranged in a columnar architecture of the edge-connected neuron array, the synapse matrix as well as the analog on-chip Capacitive Memory (Capmem) cells for the storage of tunable neuron voltage and current biases. The right half shows the digital part comprising of the SIMD plasticity processor and digital control logic. In this prototype version, the ANC contains 32 columns, each containing 32 synapse rows relaying current pulse events to a single neuron per column. Each neuron is configured by 14 current and 4 voltage biases, all of which are individually tunable per neuron as well as a 15-bit digital configuration bus. One global voltage bias is common to all neurons and provided by an additional global Capmem column, see Fig. 1a. The Capmem voltage and current cells have 10-bit resolution and provide the neuron with tunable biases of 15 nA to 1 µA (current cells) and 200 mV to 1.8 V (voltage cells). The average drift of the capacitive storage is less than 1 LSB/20 ms and the cells are periodically refreshed at a rate of 1–2 kHz to minimize its effect [27]. The columnar arrangement of neu- rons, synapses and dedicated tunable biases make the system a highly configurable non-von Neumann architecture. Data transfer between the digital backend and various components of ANC is carried out in the form of data packets via the On-chip Multi-master Non-bursting Interface Bus-fabric (OMNIBUS) [24] (blue connections in Fig. 1a). The packets 2 encode synaptic addresses, synapse matrix's row enables and the digital configuration of neurons and the Capmem. During network operation, incoming events reach the synapses via OMNIBUS containing target synapse addresses and row-wise enables. If a packet's synapse address matches the target synapse address, the target enables a 6-bit current-mode Digital to Analog Converter (DAC) for a short duration, essentially relaying a current pulse event to the neuron on either of the two dedicated lines per neuron column (2×32 in total) - one for excitatory events, another for inhibitory events. The digital input code of the 6-bit DAC inside each synapse corresponds to the synaptic weight. This is further highlighted in the left half of Fig. 1b. The digital output events evoked by each neuron are routed off-chip to the FPGA via a Serializer/Deserializer (SerDes), from where they are re-routed back into the synapse array. To provide a debug interface, the entire neuron array is connected to the chip pads by two dedicated pins, labeled Istim and VreadOut (described in Sec. IV-D). The system performs hybrid learning using parallel ana- log processing in the synapses together with arbitrarily pro- grammable learning rules in the custom SIMD plasticity pro- cessor [23]. The synapses feature a dedicated implementation of Spike-timing-dependent plasticity (STDP) [29], where they measure the temporal correlation of pre– and post-synaptic events. The post-synaptic events from all neurons are therefore routed to the synapses. To facilitate STDP, the temporal corre- lation of input/output events (pre-post or post-pre) is stored as analog voltages inside each synapse. These analog voltages are digitized by an Analog to Digital Converters (ADCs) before being read by the plasticity processor [23]. The processor can then alter the the synaptic weights (stored in SRAM) in accordance with the implemented rule, thereby performing online learning. The processor is a 32-bit implementation based on power instruction-set architecture with a custom SIMD unit. It pro- cesses 128-bit wide vectors of either eight or sixteen elements using vector slices. The current implementation features a sin- gle slice that processes 16 columns in parallel and iterates twice for 32 columns. All synapse rows are processed sequentially. The processor can be used for implementing arbitrary learning rules that make use of the available observables and modify the structure and parameters of a running network. It can additionally be used to perform local calibration algorithms that can be executed in parallel on large multi-chip systems. Within the described architecture, the total area of all ADC channels, the neurons and the number of vector slices in the processor scale linearly with the number of columns. I I I . D E S I G N M E T H O D O L O G Y When it comes to neuron design, every large scale system adheres to certain parameters governed mainly by the system specifications and target applications. The design phase of this accelerated neuron circuit in general took the following pre- considerations: A. Neuron Model Our choice of neuron model comes from the trade-off between a design that encapsulates most of the biological 3 Fig. 1: (a) Architecture of the HICANN-DLS prototype chip and the measurement system. (b) The full circuit schematic of a single integrated neuron. (c) The schematic of the resistor used inside the synaptic input. (d) The architecture of the excitatory synaptic input (swapped terminals for inhibitory synaptic input). (e) The schematic of the spike pulse generator and reset circuit (implements refractory period duration). (f) The two-stage opamp schematic used inside the read-out buffer. (g) The schematic of the source-degenerated Operational Transconductance Amplifier (OTA) (used inside synaptic input and leak). computational power versus the silicon area and design com- plexity. This entails that the emulated neuron models should be rich enough to replicate most computational studies, as well as sufficiently simple to integrate enough neurons for small functional networks on a single die. This is a good compromise to envision large-scale systems, also endorsed by other architectures, e.g., [30]. Further, Integrate and Fire (I&F) models with refractoriness and adaptation are known to reproduce the biological traces with good accuracy [31], [32]. Our models of choice are therefore low dimensional 4 mode is featured that evokes a single output spike per input synaptic event. The aforementioned digitally switchable terms are realized by the transmission gate switches Sn in Fig. 1b. An integrated voltage buffer reads out the membrane potential as well as the activity on both shared synaptic input lines within each neuron circuit. D. Power and Area Power reduction for large-scale integration is made possible at two levels: First, the Capmem in this chip prototype can provide bias currents as low as 15 nA. This saves significant power consumption, since the circuits shift to moderate or weak inversion. Further, the unused sub-circuits are power-gated. The sub-circuit design has been optimized for area/power by utilizing MOS gate capacitors, usage of thin and thick-oxide transistors and dual voltage supply options, wherever possible. As a result, the single neuron circuit consumes approximately 10 µW power (14.4 µW if both synaptic inputs are used). E. Calibration A major goal of the presented design was the possibility to calibrate individual sub-circuits as well as the entire neuron. The circuit has therefore been verified pre-silicon with focus on mismatch compensation using the available Monte Carlo models. Post-silicon, all 32 neurons have been calibrated for mismatch across multiple dies, the respective results are shown in Sec. IV. Utilizing these datasets, we can provide a mapping between LIF model quantities and circuit-level parameters by using fractional polynomial fits for the individual sub-circuits. I V. C I R C U I T D E S I G N A N D M E A S U R E M E N T S The neuron schematic designed for the current prototype chip is shown in Fig. 1b connected to a single column of the synapse matrix. Starting from the left, it shows a single synapse column from which two output lines emanate - one carry- ing excitatory pre-synaptic events and another inhibitory pre- synaptic events. Each pre-synaptic event enables a 6-bit DAC for a configurable duration of 10 ns–320 ns, which outputs a current pulse on either of the two synaptic lines. At the neuron side, these current pulses are received by the synaptic input circuits (one excitatory, one inhibitory). These circuits model the synaptic dynamics with an exponential kernel inside each neuron circuit (Fig. 1d). The latter integrate current from both synaptic inputs onto a capacitor Cmem that models the neuronal membrane. The membrane can discharge itself through a separate Leak circuit towards the resting potential Vleak. Once the voltage reaches a threshold Vthresh, the circuit SpikeGen triggers a digital pulse event (see fire in Fig. 1b,e) and brings the membrane back to a reset potential Vreset via a Reset module. This circuit also adds the respective refractory period, essentially by clamping the membrane to Vreset for a duration governed by the time it takes to toggle the inverter again. An Analog IO block reads out debug voltages and injects (test) stimulation current onto the membrane. The neuron schematic also highlights the digitally controlled interconnecting switches labeled S0−11. These are realized as transmission gates and are threshold-based I&F models such as [33], [34]. Therefore, in this initial prototype, we implement the LIF model as a base computational element. The subthreshold response of the LIF model can be defined by Kirchhoff's current law for the conservation of charge: Cmem dVmem dt = −gleak · (Vmem − Vleak) + I (1) and if Vmem ≥ Vthresh, Vmem → Vreset (2) where Vmem is the potential across the neuronal membrane, Cmem is the membrane capacitance, Vreset and Vthresh are well defined reset and threshold potentials, Vleak models the leak potential and gleak is the leak conductance. I is the sum of external current (Istim) and excitatory (Isyn,exc) and inhibitory (Isyn,inh) synaptic currents. The synaptic inputs integrating these currents are exponentially decaying current-based inputs, whose time course can be defined as: i ) (3) τsyn (cid:19)Θ(t − tf fi wie−(cid:18) t−t Isyn(t) =Xi Xf where wi is the weight of the synapse connecting a pre- synaptic neuron to a post-synaptic neuron, tf i denotes fth spike at a synapse i, Θ(x) is the Heaviside step function and τsyn is the synaptic time constant. B. Specifications and Parameter Range A selected set of computational modeling studies [35]–[46] defines the tuning range and target specifications of the individ- ual sub-circuits and eventually the neuron model parameters. Table I summarizes these identified ranges for membrane and synaptic time constants (τmem, τsyn) as well as the refractory period (τrefr) in biological real-time. The presented hardware Parameter Min. [ms] Max. [ms] τmem τsyn τrefr 50 100 10 7 1 0 TABLE I: Target specifications of the neuron model parameters. implementation uses current-based synaptic input. Because the reference studies employ conductance-based synapse models, an increased leak conductance by a factor of five is required to be able to emulate the reduction of the effective mem- brane time constant that is associated with high conductance states [47], [48]. C. Debug and Testability The neuron circuit provides the possibility to debug individ- ual sub-circuits or disconnect individual terms for verification and calibration. For example, a digitally configurable bypass meant to test, debug and switch-off individual terms whenever needed. Each of the incoming input events to either excitatory or inhibitory synaptic inputs can be made to directly trigger output events (see fireout in Fig. 1b), bypassing the current integration on the neuron membrane. Inverters with shifted trip-points driven directly by the voltage drop on the input synaptic lines make this possible - they are enabled by digital inputs S9,10 and tri-state inverters at the output. This "bypass mode" is useful for testing certain modules of the system (e.g. event routing) without relying on configured neurons. It is disabled during nominal neuron operation. Fig. 2: (a) Chip micrograph. (b) Prototype measurement sys- tem. The integrated array of 32 neuron circuits occupies 200 µm × 376 µm of die area. Each neuron in turn occupies 200 µm × 11.76 µm. The described prototype chip has an area of 1.9 × 1.9 mm2 and is fabricated in a low-K 1P9M 65-nm low-power digital CMOS process. The bonded die and the measurement setup are shown in Fig. 2. All measured analog data presented within this work has been acquired using a Keithley 2635B Sourcemeter for static current measurements and a LeCroy Wavesurfer 44Xs digital oscilloscope for dy- namic signals. The chip measurement framework comprises of an on-board Xilinx Spartan-6 FPGA system that takes command packets via a USB interface. The chip and all neuron parameters are directly programmable via a C++/Python based software environment. The neuron block specifications are summarized in Table VII. Before demonstrating the neuron operation, we describe the design and measurements of each individual sub-circuit. A. Synaptic Input The synaptic input circuit provides the exponential synap- tic dynamics to incoming events inside each neuron circuit (Eq. III-A). The circuit integrates short current pulses (synaptic events) arriving from 32 synapse circuits in a single column onto an RC integrator before converting them into an equiv- alent current with the help of a linear transconductor. The architecture is shown in Fig. 1d. The current pulse events from 32 synapse circuits are received on the shared line labeled Isyn,exc in Fig. 1b and shown as an input terminal in Fig. 1d. Each incoming input pulse event drops the voltage 5 on this line (nominally at 1.2 V), which is then recovered with the synaptic time constant τsyn. This voltage drop is proportional to the strength of the incoming pulse event. The synaptic time constant is varied by the tunable resistor Rsyn, while the integration capacitor Csyn is fixed at 1 pF. In the current prototype, this capacitance is mostly contributed by a metal capacitor, but it will eventually be realized from the line parasitics alone - as the number of input synapses will increase in the scaled-up final chip. The second OTA input Vsyn is kept at approximately 1.2 V, the precise value can be altered for calibration purposes. The realized amplifiers can cancel the effect of input offset voltage at its output using a tunable bias current IbiasOff. 1) Transconductor: The designed transconductor is a source- degenerated OTA architecture, whose output current has a linear dependence on input differential voltage within a limited range, such that Iout = Gm(Vin+ − Vin-). Where Gm is the OTA transconductance. The OTA's main bias current (labeled Ibias) helps to vary this transconductance, whereas a second bias (labeled IbiasSd) adjusts the proper biasing point of the source degeneration pair. The OTA architecture is shown in Fig. 1g. Transistors M1-M4, and M5-M8 form cascode current mirror loads, whereas M9,10 form the input pair. Transistors M11,12 are a source degeneration pair, acting therefore as degenerating resistors to lower the gain. Their bias points are tunable externally using a separate bias IbiasSd. Note the presence of two separate stages at the output: First, the output stage current multiplier formed by transistors M2h,4h and M14h,16h - it provides extra output current in parallel to the existing output stage formed by M2,4 and M14,16. A digital switch labeled enhiCon enables this current multiplication. This is useful when realizing a higher conductance - for example in the leak term to ensure short membrane time constants. The leak term (see Sec. IV-C) also uses this OTA and the output current multiplier is implemented only there. Secondly, another parallel circuit - a current mirror formed by M19,20 and M21,22 is used as an offset calibration circuit at the OTA output. Ioffset is then the calibration current that is set equal to the residual offset current, caused for example as a result of input offset voltage between the OTA terminals. The current mirror steals this current from the output stage of the OTA that directly sends current to the membrane. 2) Synaptic Resistor: The synaptic resistor Rsyn is a tunable resistor designed using bulk-drain connected devices [49]–[51] and shown in Fig. 1c. The architecture of the resistor uses a series of four pmos bulk-drain connected devices (labeled M1 to M4) between the two terminals VinP and VinN. Each pmos device connects its bulk terminal to its drain instead of the nominal bulk configuration. They provide a linear increase in drain current with increasing source-drain voltage. Series stacked devices are used to reduce the source-drain potential well below the threshold voltage. The total voltage drop across the two resistor terminals can be as large as a couple of hundred millivolts for the case of larger incoming events. For example, a 10 µA event (if 4 ns long), can drop the synaptic input line Isyn,exc/inh to 1 V, creating a drop of 200 mV across the terminals. The resistor therefore provides almost linear operation within this range. The pull-up nature of this synaptic (k) 2 −500 Synaptic Offset Current [nA] 0 500 −2 6 (d) Chips 20, 24 (e) 1.0 50 Synaptic Time Constant [µs] 1.5 10 30 (j) Chip 20 25 20 15 10 05 14 12 10 02468 Counts Counts (c) Chips 20, 22, 24 50 10 1 TimeConstant[µs] (a) (b) Chip 20 0 75 150 0 75 150 ∆V [mV] 0 100 200 300 400 500 Resistor Bias Current [nA] (h) Chip 22, N15 (i) 500 400 300 200 100 0 Counts (f) Chips 20, 22, 24 (g) 10 20 30 10 20 30 Synaptic Time Constant [µs] 10 20 30 10 20 30 Synaptic Time Constant [µs] 60 50 40 30 20 10 0 Current[nA] 180 160 140 120 100 80 60 40 20 0 Counts Fig. 3: (a) Variation of the characteristic curves for the 64 synaptic input resistors of a single chip, with bias set to 80 nA (pre-calibrated). (b) Post-calibration resistor curves with time constant of all synaptic inputs set to equal the mean of those in (a). (c) Tuning the synaptic time constants of 192 synaptic input circuits by varying the resistor bias current (d,e) Distribution of the minimum and maximum range of the achieved synaptic time constants. (f) Pre-calibration distribution of synaptic time constants with a mean of 5 µs, 10 µs and 20 µs. (g) Post calibration: the target value of (f) is processed through individual polynomial fits. (h) Trial-to-trial variation of synaptic time constants for programmed values of 5 µs, 10 µs and 25 µs. (i) Variation in synaptic time constants measured from a train of 500 continuous input events. (j) Pre-calibrated synaptic output offset current with Vsyn at 1.2 V for all OTAs. (k) Residual synaptic current after individually adjusting Vsyn and IbiasOff. input resistor ensures that source potentials are always higher than the drains. This prevents the two terminals from getting swapped and possibly consume higher currents, since it is then a nominal pmos configuration (bulk connected to source terminal). A single bias current Ibias tunes the gate voltage Device M1,2,3,4 M1c,2c,3c,4c M1b,2b,3b,4b M1a,2a,3a,4a M5,6 Width/Length [µm/µm] 0.15/1.0 0.15/5.0 0.15/0.5 0.15/0.5 0.75/0.5 TABLE II: The device dimensions of the bulk-drain connected synaptic resistor. of all subsequent bulk-drain connected devices via cascode current mirrors, labeled M5,6 and MXa–MXb, where X denotes the respective devices of each pmos stage. The devices marked MXc help tune the transistor bias points. Table II summarizes the transistor dimensions for the designed resistor. Notice that in order to provide larger resistance the channel lengths of the bulk-drain connected devices M1−4 are relatively long. We have measured synaptic input circuits on up to three different dies. The characteristic curves of the resistor within the desired voltage range are shown in Fig. 4. The figure plots the current vs. the potential difference by varying the resistor's bias (Ibias in Fig. 1c). The equivalent swept Ibias and the resulting resistances determined from a linear fit are shown. Note that at very low bias currents the resistance is very large (bottom-blue curve) and the resistance values are inclined for an exponential increase as a function of bias current. Notice the presence of finite voltage offsets, as the plotted traces cut the zero-nA line at a potential difference of about 15 mV across the terminals. This is due to supply drop and can be corrected for by tuning the input Vsyn (second OTA terminal) during operation. We further show our measurement results from all 64 re- sistors (both synaptic inputs), configured for the same mid- range resistance of about 4 MΩ in the desired voltage range in Fig. 3a. The traces show triode curves for typical MOS devices, since the bulk-drain connected devices are in fact biased in linear region to adjust the resistive range. The variation in the traces are due to device mismatch, largely contributed from the biasing stages (MXa,Xb,Xc) that help set the bias points of each individual device. This is minimized in Fig. 3b after applying the calibration described in Sec. IV-A4. 7 variation of 500 incoming events for the aforementioned three configurations. B. Spike Generator and Reset This circuit is responsible for evoking a digital output event as an indication of spike occurrence, once the membrane reaches a specified voltage threshold Vthresh. It also resets the membrane voltage to a fixed reset potential Vreset and adds a refractory period τrefr before the membrane starts integrating again. The circuit consists of a comparator whose output stage is reset after a finite delay via a feedback loop, as shown in Fig 1e. Once the membrane Vmem reaches the threshold Vthresh, the comparator outputs a logic 'high' (VOH). After a finite time delay tdelay, the comparator output stage is pulled down to logic 'low' (VOL). This essentially generates a digital voltage pulse signal fire. The fire pulse marks the spike occurrence and is routed as a single event through to the digital backend. Furthermore, it also clamps the membrane Vmem to Vreset by toggling the switch Srst (see Fig 1e). This happens as the voltage over the capacitor Crefr (which constantly integrates a tunable bias current Irefr) is reset by the incoming pulse. This initiates the neuron's refractory period (τrefr) that lasts until the voltage across the capacitor Crefr triggers the inverter again to disconnect the switch Srst (Fig 1e). This marks the end of refractory period and the membrane Vmem is then free to integrate again. The delay-cell designed for the spike pulse generator (la- beled tdelay) is a tunable current-starved delay element [52] typically set to 100 ns. The comparator is a simple two-stage architecture similar to an uncompensated two-stage op-amp that provides high gain to drive outputs to the desired output levels (VOH or VOL). To realize long refractory periods without a large Crefr, the bias current Irefr from the respective Capmem cell is divided 10 times (fixed), reducing the effective bias currents to 1.5 nA (min.) and 100 nA (max.). Our measurements of two dies (32 circuits each) suggest that the refractory period can be tuned successfully over a range from 1.11 µs up to about 137 µs, as shown in Fig. 5a,b (1σ quantile, see also Table III). The variation in longer time constants is relatively large due to very small input bias current (1.5 nA) being integrated on Crefr. The circuit has been designed for much longer refractory periods than what is known through biology to cater for artificial models such as [53]. 1) Calibration: Since the duration of the refractory period is solely determined by the current Irefr (Fig. 1e), this variation is catered for by measuring τrefr as a function of Irefr and applying a second order fractional polynomial fit for each neu- ron. In Fig. 5c,d, the distributions of three different refractory periods are shown pre- and post-calibration. Notice the larger spread towards smaller bias currents and therefore higher time constants, which is significantly reduced after calibration (c.f. Tab. IV). Fig. 5e shows the calibrated membrane potential of all 32 neurons on a single die after spike occurrence. All neurons are configured to a reset potential of 200 mV, a spike threshold of 1.2 V, a resting potential of 1 V and - in order to produce steep, comparable edges - to a short membrane time Bias, Resistance 15nA, 14.66MΩ 130nA, 3.64MΩ 240nA, 2.36MΩ 355nA, 1.82MΩ 465nA, 1.52MΩ 580nA, 1.32MΩ 695nA, 1.19MΩ 805nA, 1.08MΩ 920nA, 1.00MΩ 200 150 100 50 0 Current[nA] 0.00 0.05 0.15 0.20 0.10 ∆V [V] Fig. 4: Tunability of the synaptic resistor. The current flowing through Rsyn is measured as a function of the voltage drop across the resistor terminals, as the resistor bias is swept. 3) Full Circuit: Finally, the measured results demonstrating the range of possible synaptic time constants τsyn are shown in Fig. 3c. The data has been acquired by measuring three separate dies, each having 64 synaptic input circuits. The plot shows a range of available time constants as well as their mean value as the bias of the tunable resistor is swept from 0 to 0.5 µA. The plot shows a non-linear tuning curve which steepens substantially in the low current regime. It corresponds to the tuning behavior of synaptic resistor shown in Fig. 4. The variation in different achievable time constants, hence is larger when we tune longer synaptic time constants, compared to shorter ones. This is also evident from the histogram plot of Fig. 3d,e, where a distribution of both the shortest and longest synaptic time constants (bias settings of 1 µA and 15 nA) are plotted for 128 samples. Given these distributions, the range of available synaptic time constants is between 1.24 µs and 20.5 µs for one-sigma single sided quantiles (see Table III). 4) Calibration: The input-referred offset of the synaptic input OTA, if not canceled, can make the neuron membrane integrate unwanted output offset current. To calibrate this, we use the OTA reference voltage Vsyn as well as the offset compensating bias IbiasOff, both of which are locally tunable parameters. Using Vsyn we cancel the output offset to a first approximation. The residual offset is trimmed by sweeping IbiasOff and searching the root of a linear fit. Fig. 3j,k shows the measurement of all 64 synaptic inputs' output offset currents on a single die pre- and post-calibration. The post-calibration residual current is below 5 nA for all samples, which corre- sponds to approximately 20 mV deviation in the membrane resting potential at a membrane time constant of 10 µs. The curves shown in Fig. 3c are fitted with second order fractional polynomials, which reduces the relative spread of individual time constants. Fig. 3f,g shows the spread of three different synaptic time constants pre- and post-calibration from three dif- ferent dies. The standard deviation for these datasets is depicted in Tab. IV. Finally, we demonstrate the trial-to-trial variation for three different time constants within a single neuron. In Fig. 3h, the mean time constant of 500 incoming synaptic events is measured after reprogramming from a different time constant. Most samples are clearly restricted to the same bin. Fig. 3i on the other hand shows the synaptic time constant 8 40 (j) 0 10 20 30 40 10 20 30 Refractory Time [µs] 0 20 Time [µs] 1.0 0.5 0.0 Membrane[V] (d) (e) (c) Chips 22, 24 Chips 20, 22, 24 75 50 25 0 Counts (h) 100 (i) (g) Chips 20, 24 70 60 50 40 30 20 10 0 Counts 20 10 Counts 0 0.3 0.4 25 50 75 Membrane Time Constant [µs] 20 0 Membrane Time Constant [µs] 40 0 10 20 (b) Chips 20, 24 (a) 20 15 10 Counts 1.5 1.0 Refractory Time [µs] 150 200 B: 280, SD: 70 B: 280, SD: 90 B: 280, SD: 140 B: 375, SD: 465 B: 515, SD: 465 B: 655, SD: 465 B: Ibias [nA] SD: IbiasSd [nA] 0.5 (f) 05 012 0.4 1.2 0.0 Membrane Voltage [V] 0.8 −1 −2 Current[µA] Fig. 5: (a,b) The distribution of minimum and maximum achievable refractory time periods. (c,d) Pre- and post-calibration results of refractory period circuits, set for a mean of 5 µs, 15 µs and 25 µs respectively. (e) Calibrated membrane traces showing a post-spike response of different on-chip neurons with refractory times of 10 µs and 25 µs. (f) Output current of the leak term as the membrane potential is swept with the resting potential fixed at 0.52 V. (g,h) Distribution of the minimum and maximum achievable membrane time constants. (i,j) Pre- and post-calibration distribution for τmem of 1 µs, 10 µs and 20 µs. constant of 250 ns. The figure shows two traces per neuron, one programmed for a refractory time of 10 µs and the other for 25 µs. Note how all neurons reach the resting potential with little remaining spread in the time constant. The visible variations of the resting potential are due to residual offset currents from the synaptic inputs (c.f. Sec. IV-A4), as well at least 31 mV of input offsets of the leak OTAs (estimated through simulation using Monte Carlo device models). C. Membrane Leak and Capacitance The membrane conductance has been realized using a linear transconductor whose output terminal is fed back to its negative input as shown in Fig. 1b. In this configuration, the total conductance across the input and output terminals equals the transconductance Gm. The usage of linear transconductors to realize conductance is already well known in neuromorphic architectures [5], its architecture is described in Sec. IV-A1. A two-bit configurable gate oxide capacitor provides a total of 2.36 pF membrane capacitance. This is implemented in multiples of 590 fF for layout symmetry, switchable between 2.36 pF, 1.77 pF or 590 fF via S7-8 (see Fig. 1b). To minimize distortions introduced by the inherently non-linear behavior of MOS capacitors, the respective transistors were chosen to be thick-oxide devices. Within the membrane's full dynamic range of up to 1.2 V, the bulk-gate drop can therefore be kept above 1.3 V and the device in inversion region where the gate capacitance is constant. Parameter 1 min (1σ, 3σ)[µs] max (1σ, 3σ)[µs] τrefr τsyn τmem 2 1.11, 1.24 1.24, 1.41 0.35, 0.41 137.5, 104.5 20.5, 13.6 16.6, 11.3 1 Single-sided 1σ (3σ) quantiles of 84.13% (99.86%) 2 Measured using Cmem = 2.36 pF; The min./max. τmem (1σ, 3σ) for Cmem = 570 fF is estimated to be 0.08, 0.10 µs and 4.11, 2.82 µs respectively TABLE III: Measured range of time constants from data acquired across two dies. In Fig. 5g,h, we show the distribution of achieved membrane time constants with the largest selectable membrane capaci- tance. The distribution highlights the minimum and maximum achievable time constants (τmem) from a total of 64 leak circuits on two different dies. It is shown that the neuron can set time constants, as small as 0.35 µs and up to 16.6 µs (one-sigma quantiles) and can be further reduced by selecting different capacitor configurations (c.f. Table III). Fig. 1b shows how the leak term uses the transconductor in a feedback configuration, as opposed to the synaptic input where the OTA is in open loop. In feedback configuration, the OTA transconductance Gm is the overall conductance of the leak term, defining membrane time constant as τmem = Cmem/Gm. target 5 µs 15 µs 25 µs 5 µs 10 µs 20 µs 1 µs 10 µs 20 µs pre-calibration 0.50 µs (9.95%) 1.84 µs (12.3%) 4.88 µs (18.9%) 0.75 µs (15.2%) 2.49 µs (20.7%) 4.01 µs (19.9%) 0.19 µs (19.3%) 3.14 µs (40.9%) 8.34 µs (42.3%) post-calibration 0.14 µs (2.83%) 0.12 µs (0.77%) 0.51 µs (2.06%) 0.14 µs (2.63%) 0.18 µs (1.75%) 1.12 µs (5.56%) 0.03 µs (3.03%) 0.46 µs (4.53%) 2.27 µs (11.5%) τrefr τsyn τmem TABLE IV: Pre- and post calibration standard deviations for the histograms shown in Fig. 3 and Fig. 5. Fig. 5f shows six measured traces plotting the output current of the leak term as a function of the membrane voltage. Three of them sweep the main bias current Ibias keeping IbiasSd constant, while the remaining three sweep IbiasSd, while Ibias is held constant. The traces highlight how lowering the source degeneration bias linearizes the response and allows for a wider range of membrane voltages. This is because the degeneration resistors, implemented as transistors M11, M12, are shifted from linear region to the saturation region of the weak inversion regime, where their on-resistance Rs is as large as few MΩ at low IbiasSd. This effectively linearizes the curve as the OTA's transconductance Gm approximates to 1/Rs. 1) Calibration: The membrane time constant can be tuned by three parameters: A two-bit variable membrane capacitor and the two biases (Ibias, IbiasSd) of the leak OTA. Despite the low gain curves and consequently wide range for lower bias values of IbiasSd (Fig. 5f), tuning is done using the former two parameters. The circuit achieves membrane time constants between 0.35 µs and 16.6 µs within a linear range of about 230 mV. The reason for not relying on IbiasSd is twofolds - first, low values of IbiasSd contribute a large input offset, causing an output residual current, that is not calibrated for by tuning Vleak. Secondly, at lower values of IbiasSd, overall transconduc- tance Gm contributed by the OTA is a non-monotonic function of Ibias. To avoid both effects, we calibrate the membrane time constant by adjusting the OTA's bias current (Ibias) and therefore trade linear range for robustness. We measure the relation between τmem and Ibias at fixed IbiasSd of 1 µA for a voltage decay from 600 mV towards a resting potential of 400 mV. The resulting curves are fitted by second degree fractional polynomials. Pre- and post-calibration variations of τmem are depicted in Fig. 5i,j for targets of 1 µs, 10 µs and 20 µs, the related standard deviations are shown in Tab. IV. It may be noted that not all neurons can achieve the maximum time time constant of 20 µs. These neurons are visible as outliers to the left of the depicted distribution. D. Analog Input/Output We can inject constant current into the membrane, hold it to a fixed potential or read out debug voltages via the analog I/O sub-circuit (see Istim and VreadOut in Fig. 1b). 9 Output range Load –1-dB bandwidth Gain Power Input offset Slew rate Compensation cap. 0.1–2.1 V ≥ 16 pF k10 MΩa 1.15 MHzb 62.6 dBc 127 µWc 14 mV 2 V/µs 600 fF TABLE V: Measured specifications of the read-out buffer. aEstimated off-chip load bMeasured after the LPF formed by output Tx-gate and parasitic load cSimulation only Device M1a,1b,2a,2b M3,4 M6, M7 M12, M5 M9,11 M8,10 Width/Length [µm/µm] 2.4/0.28 0.8/0.38 4.0/0.38 & 4.0/0.39 0.15/0.75 & 1.6/0.39 3.6/0.38 0.6/0.38 TABLE VI: The device sizes of the read-out amplifier. The monitored output voltage is multiplexed from among the membrane potential or incoming synaptic activity at any of the two synaptic input lines. The Opamp designed for the buffer is a two-stage amplifier with split-length indirect compensation [54], [55] and is shown in Fig. 1f. The buffer has a –1-dB bandwidth of 1.15 MHz when driving large off-chip loads at a maximum power consumption of 127 µW. The close- loop opamp bandwidth is filtered by the output low-pass filter formed by transmission gate and the off-chip load. In the current design, each neuron buffer directly drives the shared read-out line that terminates at the output pad. In comparison to Miller compensation, the use of indirect compensation ensures smaller area, higher bandwidth, and reduced power consump- tion [55]–[57]. During nominal neuron operation, the amplifier is switched off, the stated power is only consumed during analog read-out (therefore not reported in Sec. III-D). The measured amplifier specifications are summarized in Table V and the transistor dimensions used in this design are listed in Table VI. E. Full Circuit Characterization Here we describe the measurements and performance of the complete neuron circuit. Fig. 6 shows equidistant incoming events at the excitatory synaptic input line (bottom trace) and the resulting membrane trace (top). The reset potential is set to 0.2 V, the leak potential to 0.58 V and the spiking threshold to 1.2 V. The input synaptic lines (bottom trace) have a constant 1.2 V potential unless an input event arrives. At event arrival, this level drops and is recovered with a time constant τsyn. Each incoming event increases the membrane potential, eventually 10 0 20 40 80 60 Time [µs] 100 120 140 Fig. 6: Measured results - bottom: synaptic input potential indicating incoming events as evaluated on Isyn,exc line, top: resulting membrane response to the synaptic events. 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 1.2 1.1 1.0 Membrane[V] Input[V] Leaky I&F 32 11.76 × 200 µm2 2.5/1.2 V 65-nm CMOS Neuron model No. of neurons Area of single neuron Voltage supply Process Speed-up (acceleration) factor ×1000 Global parameters Local (individual) parameters Configurability Membrane capacitor (max.) Input synaptic event (max.) 1 voltage bias 1 18 (14 I-bias, 4 V-bias)1 15-bit digital bus 2.36 pF (2-bit config.) 10 µA, 4 ns pulses2 1 available from on-chip tunable capacitive memory 2 amplitude and length of each current pulse emitted by cells the synapse circuit b) Software Neuron: Membrane Potential c) Hardware Neuron: Synaptic Input d) Hardware Neuron: Membrane Potential 200 100 0 Istim[nA] 900 800 700 600 Vmem[mV] 120 60 0 ∆Vsyn[mV] 900 800 700 600 Vmem[mV] 0.0 0.2 0.4 0.6 Time [ms] 0.8 1.0 Fig. 7: A one-to-one comparison of the emulated neuron vs. software simulation. F . Power Consumption In order to estimate the chip prototype's power consumption, the total current drawn by the chip during operation of the network described in Sec. V has been measured. From this measurement, we infer a mean full-chip power consumption of (48.62± 0.04) mW. This figure includes prototype-specific consumers - especially the 2.5 V full-swing IO channels that communicate each spike off-chip to the controlling FPGA. This IO power is expected to scale sublinearly with increasing chip size. evoking a spike, followed by a membrane reset to initiate the refractory period duration. We further demonstrate the neuron parameter set from a typical modeling study [43] and compare the response proper- ties of our emulated on-chip neuron, with that of a simulated accelerated LIF model neuron (Eq. III-A). We use the Brian spiking neural network simulator [58] to simulate the ideal LIF neuron. The on-chip neuron circuit here is uncalibrated and tuned manually such that the most pronounced non-ideal effects are canceled for. This includes the output residual current of synaptic input due to input offset voltage between the OTA terminals. Further, time-constants are fine-tuned and synaptic weights are adjusted to match the spike times of the software simulation. The synaptic and membrane time constants (τsyn, τmem) are set as 1.5 µs and 10 µs, with a refractory period τrefr of 2 µs (×1000 speed-up from biological real-time). The threshold is set to 870 mV and the leak and reset potentials are 600 mV. A random spike train stimulus containing 32× 20 events (each synapse relaying 20 pre-synaptic events) is in- jected into both neurons and their corresponding membrane traces are plotted in Fig. 7. Fig. 7a shows the input current as a result of random input stimulus of the software neuron, while Fig. 7b shows its corresponding membrane voltage. In the hardware domain, we monitor the on-chip voltage on the excitatory synaptic input line, as its voltage drop (defined as ∆Vsyn := VSyn − Vsyn,exc, see Fig. 1d) is in proportion to the charge of each incoming synaptic event. A larger voltage drop corresponds here to higher input synaptic current (Fig. 7c). Fig. 7d shows the measured membrane voltage of the on-chip neuron. Comparing the dynamics of both membrane traces, one finds the RMS of their common-mode corrected difference to be 45 mV. We see that, when presented with the same input, the on-chip neuron has similar spike-times and the membrane voltage follows the time course of the software simulation. TABLE VII: A summary of neuron array specifications. a) Software Neuron: Synaptic Input 11 0 20 40 60 80 100 120 Stimulation [kHz] 0 20 40 60 80 100 120 Stimulation [kHz] 120 100 80 60 40 20 0 (c) Response[kHz] 40 30 20 10 0 (b) Response[kHz] (a) Fig. 8: (a) Schematic diagram of the implemented network topology. All neurons in the network form self- and direct-neighbor excitatory projections, as well as to central inhibitory neurons. Both inhibitory neurons in turn inhibit the ring neurons and do not make connections between themselves. (b) Network response for two competing input stimuli r1 and r2 before the network routing has been established: One neuron receives Poisson noise with an average rate of r1 = 50 kHz, another neuron receives Poisson noise with variable frequency between r1 = 0 kHz and r1 = 120 kHz. The mean firing rate of these neurons and their next neighbors is depicted as an average over 10 trials with different random seeds for the input stimulus. (c) Network response for two competing input stimuli r1 and r2 after network routing has been established: The Poisson input is as specified in Fig. 8b. An estimation of the mean energy cost of a single spike has been compiled by comparing the aforementioned result against the power consumption of an idle chip (without stimulating spikes) with the same network configuration. Due to the greatly reduced number of expensive outgoing traffic when the system is idle, the resulting number has to be understood as an upper bound for the on-chip spike cost. We find Espike ≤ (790 ± 170) pJ for the presented network. V. A W I N N E R - TA K E - A L L N E T W O R K We demonstrate a soft winner-take-all circuit [59]–[61] using the on-chip LIF neuronal array. From an array of 32 neurons, 20 are used to form a ring of excitatory neurons, each having self-excitation, nearest-neighbor excitatory projection, and excitatory projections to two common inhibitory neurons. The inhibitory pool of two neurons in turn inhibit all neurons in the ring and do not make any projection between themselves. The resulting network topology is shown in Fig. 8a. All neurons are chosen randomly and are calibrated as described in the previous sections. To verify the network's behavior, two opposite (equidistant) neurons in the ring network are stimulated with Poisson stimuli of various frequencies between 0 kHz and 120 kHz. Fig. 8c shows the average firing rates of both competing sub-populations. A sub-population here is defined as the stimulated neuron and its two direct neighbors. The stimulus to one is fixed at 50 kHz, while to the other is being swept. The swept stimulus has an increment of 0.5 kHz between separate runs, accounting to 240 different stimuli runs within the specified range. Each experiment has been run for 100 ms. The line/highlighted region in Fig. 8c shows the average and standard deviation over 10 independent runs; each with a different random seed for the Poisson input. Due to the speed-up of the presented neuron array (compared to biological time scales), the total network emulation time of four minutes (10 trials, each consisting of 240 different stimuli `a 100 ms) corresponds to 2.8 days of biological network activity. The mean firing rate for both populations during 10 control runs with disabled spike connectivity (routing) is shown in Fig. 8b. As in Fig. 8c, the mean firing frequency of one single sub- population is plotted - as there are no connections, both neighbors are quiet and population firing rate is scaled roughly by a factor of three. V I . D I S C U S S I O N It has been demonstrated that We have presented an array of LIF neurons designed for the initial prototype of our HICANN-DLS chip. The pa- per summarized comprehensive measurement and calibration results to demonstrate the usability of the realized neuron array. the statistical spread of the time constants can be minimized significantly, once they are calibrated for device mismatch and corner variations. The transconductor architecture features an offset cancellation mechanism that reduces the residual synaptic output offset by two orders of magnitude. The usability of the circuit has been demonstrated by its low trial-to-trial variation. In comparison to synaptic time constants, the membrane time constants show wider spread, since the former was tuned via the synaptic resistor which has a better control compared to leak OTA biases. We conclude that both synaptic as well as membrane time con- stants need improvement to cover longer durations in a future revision. The circuit evokes a digital event as an output spike that also initiates the refractory period - a very broad duration has been demonstrated with a statistical spread that is widely reduced after calibration. The presented prototype is ready to i i-1 i+1 excitatory inhibitory CMOS tech. [nm] 28 Architecture Speed-up Neuron Model TrueNorth Neurogrid 180 analog subth. ×1 quad. I&F Yes digital ×1 augmented LIF No This Work 65 analog ×1000 LIF Yes Biophysical Dynamicsa Neuron Area [µm2] Local Parameters Parameter Area [µm2] 1800 No – 2352c Yes 2688d 2900b Yes – a tunable τrefr, τmem, τsyn b multiplexed 256 times per time step c including excitatory and inhibitory synaptic inputs d area of 18 analog on-chip biases per neuron for current work; not comparable to TrueNorth's 124-bit SRAM- based parameter config. or Neurogrid's shared biases TABLE VIII: An overview of neuron model specifications in large-scale neuromorphic architectures. run local calibration routines using the embedded processor, thereby enabling highly parallel calibration of multiple neurons in large-scale systems. Such local calibration algorithms are subject to future development. To verify the characteristics of the emulated physical neuron, a comparison with a software-simulated ideal LIF neuron is made: the number of spikes as well as the membrane dynamics are found to be in close correspondence. Finally, using the chip routing resources, the FPGA and system software stack, a small soft winner-take-all network is realized. A basic functionality of the circuit to two competing sub-populations. It is pertinent to note that usage of accelerated hardware reduced the network run-time (compared to biological real-times) by three orders of magnitude, enabling a platform feasible to study long-term developmental processes in the nervous system. is shown by feeding Poisson input We further compare our chip's neuron implementation with two state-of-the-art neuromorphic implementations, namely the IBM's TrueNorth system [15] and Stanford University's Neu- rogrid architecture [9]. The TrueNorth neuron is a phenomeno- logical LIF model that implements various synthetic, arithmetic and logical operations along with real-time spiking behaviors. Being a phenomenological model, the temporal dynamics of the synaptic and membrane time constants (τsyn, τmem) are not tunable implementations. Despite a 28 nm CMOS implementa- tion, the digital circuit occupies more area (2900 µm2), reduced effectively by the possibility of time-multiplexing. Alternate phenomenological implementations such as the one presented by Loihi's 14 nm architecture offer slightly worse neuron density [16]. Compared to this, Neurogrid - like BrainScaleS 12 TrueNorth Neurogrid Hi-DLS Die Area [mm2] Neurons/core [#] Neurons/timea[#] Synapses/core [#] SynOps/sec [GSOPS] 430 4k 1M 64k 58 168 64k 64k 256kb – 32 512 512k 131k 16k a emulated neurons per timescale: HICANN-DLS acceler- ates ×1000; TrueNorth multiplexes ×256 b each synapse population (4× per neuron) counted as individual synapse TABLE IX: An overview of single-chip characteristics of large- scale neuromorphic systems with scaled-up HICANN-DLS chip. Data plotted in Fig. 9. - offers biologically plausible dynamics and tunable time constants. However, all 64 k neurons in a single Neurocore share their parameters, whereas the current work has dedicated neuron parameters. Neurogrid implements subthreshold MOS dynamics, which typically allow for compact neurons due to small currents leading to less capacitor area for given time constants. Both Neurogrid and TrueNorth operate in real-time, as opposed to our accelerated approach. The neuron area occu- pied here is slightly larger than Neurogrid neuron - in which about 500 µm2 is occupied by membrane capacitor (MOScap). The neuron model we implement in the current prototype is a LIF model, compared to Neurogrid's two compartment quadratic model. The presented modular neuron architecture has been enhanced to an Adaptive-Exponential I&F (AdEx) model with multiple compartments in a recently submitted chip prototype [62]–[65]. Table VIII summarizes this comparison against the TrueNorth and Neurogrid neurons. The measured energy per output spike (Sec. IV-F) includes the off-chip routing and has been reduced by providing local (on-die) routing in the next prototype. We have not measured energy per synaptic activation in our chip which has been measured at 941 pJ and 26 pJ for implemented networks in Neurogrid and TrueNorth systems. We expect that the upper bound of (790 ± 170) pJ of energy/spike will reduce with more neurons and input synapses in the scaled-up system. The energy consumed in the stated neuromorphic architectures are orders of magnitude lower than, for example, a quad-core Xeon CPU [66] simulating a spiking neural network and still more efficient compared to the network emulated on many-core embedded systems, e.g. on the SpiNNaker system [67]. Fig. 9 presents a multi-dimensional overview of this proto- type chip with the scaled-up HICANN-DLS as well as the single-chip TrueNorth and Neurogrid implementations. The chip area, number of physical neurons and synapses are shown against peak synaptic operations per second (SOPS) and em- ulated neurons in a given time-scale. The resulting pentagons reflect the architectural preference of each implementation. Data about peak SOPS was not available for Neurogrid. The TrueNorth chip with the largest die area (430 mm2) time- muliplexes the 4096 physical neuron instances to realize the 13 system architect and conceived the initial OTA and resistor architectures, later implemented by S.A.A. The conceptual advice was given by K.M. The authors would like to extend their gratitude to Matthias Hock for support during design and measurement. R E F E R E N C E S [1] E. R. Kandel, J. H. Schwartz, and T. M. Jessell, Principles of Neural [2] [3] Science, 4th ed. New York: McGraw-Hill, 2000. J. von Neumann, "First draft of a report on the edvac," IEEE Annals of the History of Computing, vol. 15, no. 4, pp. 27–75, 1993. J. von Neumann, The computer and the brain. New Haven, CT, USA: Yale University Press, 1958. [4] K. Meier, "Special report : Can we copy the brain? - the brain as computer," IEEE Spectrum, vol. 54, no. 6, pp. 28–33, June 2017. [5] C. A. Mead, Analog VLSI and Neural Systems. Reading, MA: Addison Wesley, 1989. [6] G. E. Moore, "Cramming more components onto integrated circuits," Proceedings of the IEEE, vol. 86, no. 1, pp. 82–85, 1998. [7] G. Indiveri and S.-C. Liu, "Memory and information processing in neuromorphic systems," Proceedings of the IEEE, vol. 103, no. 8, pp. 1379–1397, 2015. [8] S. Furber, "Large-scale neuromorphic computing systems," Journal of Neural Engineering, vol. 13, no. 5, p. 051001, 2016. [9] B. V. Benjamin et al., "Neurogrid: A mixed-analog-digital multichip system for large-scale neural simulations," Proceedings of the IEEE, vol. 102, no. 5, pp. 699–716, 2014. [10] R. J. Vogelstein et al., "Dynamically reconfigurable silicon array of spiking neurons with conductance-based synapses," IEEE Transactions on Neural Networks, vol. 18, no. 1, pp. 253–265, Jan 2007. [11] Y. Wang and S.-C. Liu, "A two-dimensional configurable active silicon dendritic neuron array," IEEE Transactions on Circuits and Systems I: Regular Papers, vol. 58, no. 9, pp. 2159–2171, 2011. [12] S. Brink, S. Nease, P. Hasler, S. Ramakrishnan, R. Wunderlich, A. Basu, and B. Degnan, "A learning-enabled neuron array IC based upon transistor channel models of biological phenomena," IEEE Transactions on Biomedical Circuits and Systems, vol. 7, no. 1, pp. 71–81, 2013. [13] N. Qiao et al., "A reconfigurable on-line learning spiking neuromorphic processor comprising 256 neurons and 128k synapses," Frontiers in Neuroscience, vol. 9, p. 141, 2015. [14] E. Chicca, F. Stefanini, C. Bartolozzi, and G. Indiveri, "Neuromorphic electronic circuits for building autonomous cognitive systems," Proceed- ings of the IEEE, vol. 102, no. 9, pp. 1367–1388, Sept 2014. [15] P. A. Merolla et al., "A million spiking-neuron integrated circuit with a scalable communication network and interface," Science, vol. 345, no. 6197, pp. 668–673, 2014. [16] M. Davies et al., "Loihi: A neuromorphic manycore processor with on- chip learning," IEEE Micro, vol. 38, no. 1, pp. 82–99, January 2018. [17] H. Markram, "The human brain project," Scientific American, vol. 306, no. 6, pp. 50–55, 2012. [18] FACETS, "Fast Analog Computing with Emergent Transient States – project website," http://www.facets-project.org, 2010. [19] S. B. Furber et al., "The SpiNNaker project," Proceedings of the IEEE, vol. 102, no. 5, pp. 652–665, May 2014. [20] K. Meier, "A mixed-signal universal neuromorphic computing system," [21] [22] in Proc. IEDM, Dec 2015, pp. 4.6.1–4.6.4. J. Schemmel, D. Bruderle, A. Grubl, M. Hock, K. Meier, and S. Millner, "A wafer-scale neuromorphic hardware system for large-scale neural modeling," in Proc. ISCAS, 2010, pp. 1947–1950. J. Schemmel, "BrainScales 2: A novel architecture for analog acceler- ated neuromophic computing and hybrid plasticity," Internal document, ASIC Lab., Kirchhoff-Institute for Physics, 2017. maximum number of neurons and synapses, while Neurogrid operates in real-time and features the largest physical neuron count (65 k per Neurocore). The HICANN-DLS chip features a 6-bit STDP enabled synapse circuit, which cannot be com- pared with Neurogrid's analog-diffusor network and synapse- population circuits (which we count here as individual synapse) - and neither with TrueNorth's 1-bit crossbar synapses. For the least die area, HICANN-DLS maximizes synaptic operations per second due to the time-acceleration and possibility of high input data rates. We emphasize that Fig. 9 and its correspond- ing data given in Table IX may infer a very limited set of architectural and operational features integrated by each large- scale system. It highlights how the scaled-up chip of current prototype will feature against other single-chip building blocks. A direct comparison pertaining to varying architectural designs and system capabilities is not possible. Emulated neurons per time 7 10 1014 Synaptic operations [SOPS] This prototype HICANN-DLS TrueNorth Neurogrid Die Area [m2] 1 0 − 6 0 10 1 0 6 Physical neurons Fig. 9: Overview of the single-chip implementations in con- temporary large-scale neuromorphic architectures. V I I . C O N C L U S I O N We have presented the design, measurement, calibration and application of our mixed-signal 65-nm CMOS LIF neuronal array for the second-generation BrainScaleS platform. The neurons cover a wide tunable parameter range to act as a general-purpose element for many computational models. Hav- ing demonstrated a functional scaled-down prototype of the 65- nm chip in tight-integration with synaptic matrix and analog capacitive memory, we march on towards the development of an enhanced scaled-up HICANN-DLS chip for our large-scale bio-physically inspired neuromorphic system. A U T H O R C O N T R I B U T I O N A N D A C K N O W L E D G M E N T S.A.A. wrote the manuscript, designed the neuron circuits and implemented the on-chip array. Y.S. performed chip mea- surements, post tape-out calibration and the network demon- stration. P.M. devised the pre tape-out calibration and provided biological benchmarks. C.P. implemented the external FPGA spike routing. A.H. and A.G. implemented the digital backend and performed top-level verification. J.S. was the overall 10 9 Synapses [23] S. Friedmann, J. Schemmel et al., "Demonstrating hybrid learning in a flexible neuromorphic hardware system," IEEE Trans. BioCAS, vol. 11, no. 1, pp. 128–142, Feb 2017. [24] S. Friedmann, "A new approach to learning in neuromorphic hardware," Ph.D. dissertation, Ruprecht-Karls-Universitat Heidelberg, 2013. [25] S. A. Aamir, P. Muller, A. Hartel, J. Schemmel, and K. Meier, "A highly tunable 65-nm CMOS LIF neuron for a large scale neuromorphic system," in Proc. ESSCIRC, Sept 2016, pp. 71–74. [26] M. Hock, "Modern semiconductor technologies for neuromorphic hard- ware," Ph.D. dissertation, Ruprecht-Karls-Universitat Heidelberg, 2014. [27] M. Hock, A. Hartel, J. Schemmel, and K. Meier, "An analog dynamic memory array for neuromorphic hardware," in Proc. ECCTD, Sep. 2013, pp. 1–4. [28] C. Bartolozzi et al., Neuromorphic Systems. John Wiley and Sons, Inc., 2016. [29] H. Markram, J. Lubke, M. Frotscher, and B. Sakmann, "Regulation of synaptic efficacy by coincidence of postsynaptic aps." Science, vol. 275, pp. 213–215, 1997. J. V. Arthur and K. A. Boahen, "Silicon-neuron design: A dynamical systems approach," IEEE Transactions on Circuits and Systems I: Regular Papers, vol. 58, no. 5, pp. 1034–1043, 2011. [30] [31] R. Jolivet et al., "Integrate-and-fire models with adaptation are good enough: predicting spike times under random current injection," NIPS, vol. 18, pp. 595–602, 2006. [32] W. Gerstner, W. Kistler, R. Naud, and L. Paninski, Neuronal Dynamics. Cambridge University Press, 2014. [33] L. Lapicque, "Recherches quantitatives sur l'excitation electrique des nerfs traitee comme une polarization," Journal de Physiologie et Pathologie General, vol. 9, pp. 620–635, 1907. [34] R. Brette and W. Gerstner, "Adaptive exponential integrate-and-fire model as an effective description of neuronal activity," J. Neurophysiol., vol. 94, pp. 3637 – 3642, 2005. [35] A. Destexhe, "Self-sustained asynchronous and Up/Down states in thalamic, cortical and thalamocortical networks of nonlinear integrate-and-fire neurons." Journal of Computational Neuro- science, vol. 3, pp. 493 – 506, 2009. irregular states [36] A. Destexhe and D. Contreras, "Neuronal computations with stochastic network states," Science, vol. 314, no. 5796, pp. 85–90, 2006. [37] B. Nessler et al., "Bayesian computation emerges in generic cortical microcircuits through spike-timing-dependent plasticity," PLoS Compu- tational Biology, vol. 9, no. 4, p. e1003037, 2013. [38] T. P. Vogels and L. F. Abbott, "Signal propagation and logic gating in networks of integrate-and-fire neurons," J Neurosci, vol. 25, no. 46, pp. 10 786–95, Nov 2005. [39] G. Deco and V. K. Jirsa, "Ongoing cortical activity at rest: critical- ity, multistability, and ghost attractors," The Journal of Neuroscience, vol. 32, no. 10, pp. 3366–3375, 2012. [40] R. Naud, N. Marcille, C. Clopath, and W. Gerstner, "Firing patterns in the adaptive exponential integrate-and-fire model," Biological Cybernet- ics, vol. 99, no. 4, pp. 335–347, Nov 2008. [41] M. A. Petrovici, B. Vogginger, P. Muller, O. Breitwieser et al., "Charac- terization and compensation of network-level anomalies in mixed-signal neuromorphic modeling platforms," PloS one, vol. 9, no. 10, 2014. [42] M. A. Petrovici, I. Bytschok, J. Bill, J. Schemmel, and K. Meier, "The high-conductance state enables neural sampling in networks of LIF neurons," BMC Neuroscience, vol. 16, no. Suppl 1, p. O2, 2015. J. Kremkow, L. Perrinet, G. Masson, and A. Aertsen, "Functional consequences of correlated excitatory and inhibitory conductances in cortical networks." J Comput Neurosci, vol. 28, pp. 579–594, 2010. [43] [44] M. Pospischil, Z. Piwkowska, M. Rudolph, T. Bal, and A. Destexhe, "Calculating event-triggered average synaptic conductances from the membrane potential," J. Neurophysiology, vol. 97, p. 2544, 2007. [45] T. Masquelier and G. Deco, "Network bursting dynamics in excitatory 14 cortical neuron cultures results from the combination of different adaptive mechanism," PloS one, vol. 8, no. 10, p. e75824, 2013. [46] S. Millner, "Development of a multi-compartment neuron model emu- lation," Ph.D. dissertation, Heidelberg University, 2012. [47] D. Pare et al., "Impact of spontaneous synaptic activity on the resting properties of cat neocortical pyramidal neurons in vivo," J Neurophysiol, vol. 79, no. 3, pp. 1450–60, 1998. [48] A. Destexhe et al., "The high-conductance state of neocortical neurons in vivo," Nature Rev. Neurosci., vol. 4, pp. 739–751, 2003. [49] F. Cannillo et al., "Bulk-drain connected load for subthreshold MOS current-mode logic," Electron. Lett., vol. 43, no. 12, June 2007. [50] F. Cannillo et al., "Nanopower subthreshold MCML in submicrometer CMOS technology," IEEE Transactions on Circuits and Systems I: Regular Papers, vol. 56, no. 8, pp. 1598–1611, Aug 2009. [51] A. Tajalli, E. J. Brauer, Y. Leblebici, and E. Vittoz, "Subthreshold source- coupled logic circuits for ultra-low-power applications," IEEE J. Solid- State Circuits, vol. 43, no. 7, pp. 1699–1710, July 2008. [52] D. K. Jeong, G. Borriello, D. A. Hodges, and R. H. Katz, "Design of PLL-based clock generation circuits," IEEE J. Solid-State Circuits, vol. 22, no. 2, pp. 255–261, Apr 1987. [53] M. Petrovici, J. Bill, I. Bytschok, J. Schemmel, and K. Meier, "Stochastic inference with spiking neurons in the high-conductance state," Physical Review E, vol. 94, no. 4, October 2016. [54] V. Saxena and R. J. Baker, "Indirect compensation techniques for three- stage fully-differential op-amps," in Proc. MWCAS, Aug 2010, pp. 588– 591. [55] S. A. Aamir, P. Harikumar, and J. J. Wikner, "Frequency compensation of high-speed, low-voltage CMOS multistage amplifiers," in Interna- tional Symposium on Circuits and Systems, May 2013, pp. 381–384. split-length transistors," in Proc. MWSCAS. [56] V. Saxena and R. J. Baker, "Compensation of CMOS op-amps using IEEE, 2008, pp. 109–112. [57] S. A. Aamir, P. Angelov, and J. J. Wikner, "1.2-v analog interface for a 300-msps HD video digitizer in core 65-nm CMOS," IEEE Transactions on VLSI Systems, vol. 22, no. 4, pp. 888–898, April 2014. [59] [58] D. Goodman and R. Brette, "Brian: a simulator for spiking neural networks in Python," Front. Neuroinform., vol. 2, no. 5, 2008. J. Lazzaro, S. Ryckebusch, M. Mahowald, and C. Mead, "Winner-take- all networks of o (n) complexity," in NIPS, vol. 1, 1988, pp. 703–711. [60] R. Hahnloser et al., "Digital selection and analogue amplification coexist in a cortex-inspired silicon circuit," Nature, vol. 405, no. 6789, pp. 947– 951, 2000. [61] T. Pfeil et al., "Six networks on a universal neuromorphic computing substrate," Frontiers in Neuroscience, vol. 7, p. 11, 2013. [62] S. A. Aamir, "Mixed-signal circuit implementation of spiking neuron models," Ph.D. dissertation, Karlsruhe Institute of Technology, 2018. [63] S. A. Aamir, P. Muller, L. Kriener, G. Kiene, J. Schemmel, and K. Meier, "From LIF to AdEx neuron models: accelerated analog 65 nm CMOS implementation," in Proc. BioCAS, Oct. 2017, pp. 1–4. [64] S. A. Aamir, P. Muller, G. Kiene, L. Kriener, Y. Stradmann, A. Grubl, J. Schemmel, and K. Meier, "A mixed-signal structured AdEx neuron for accelerated neuromorphic cores," IEEE Transactions on Biomedical Circuits and Systems, 2018. J. Schemmel, L. Kriener, P. Muller, and K. Meier, "An accelerated analog neuromorphic hardware system emulating NMDA- and calcium-based non-linear dendrites," in Proc. IJCNN, May 2017, pp. 2217–2226. [65] [66] P. S. Paolucci et al., "Power, energy and speed of embedded and server multi-cores applied to distributed simulation of spiking neural networks: ARM in NVIDIA Tegra vs Intel Xeon quad-cores," CoRR, vol. abs/1505.03015, 2015. [67] E. Stromatias et al., "Scalable energy-efficient, low-latency implemen- tations of trained spiking deep belief networks on SpiNNaker," in Proc. IJCNN, July 2015, pp. 1–8.
1012.0490
2
1012
2011-01-10T16:42:46
Testing of information condensation in a model reverberating spiking neural network
[ "q-bio.NC", "cs.NE" ]
Information about external world is delivered to the brain in the form of structured in time spike trains. During further processing in higher areas, information is subjected to a certain condensation process, which results in formation of abstract conceptual images of external world, apparently, represented as certain uniform spiking activity partially independent on the input spike trains details. Possible physical mechanism of condensation at the level of individual neuron was discussed recently. In a reverberating spiking neural network, due to this mechanism the dynamics should settle down to the same uniform/periodic activity in response to a set of various inputs. Since the same periodic activity may correspond to different input spike trains, we interpret this as possible candidate for information condensation mechanism in a network. Our purpose is to test this possibility in a network model consisting of five fully connected neurons, particularly, the influence of geometric size of the network, on its ability to condense information. Dynamics of 20 spiking neural networks of different geometric sizes are modelled by means of computer simulation. Each network was propelled into reverberating dynamics by applying various initial input spike trains. We run the dynamics until it becomes periodic. The Shannon's formula is used to calculate the amount of information in any input spike train and in any periodic state found. As a result, we obtain explicit estimate of the degree of information condensation in the networks, and conclude that it depends strongly on the net's geometric size.
q-bio.NC
q-bio
Testing of information condensation in a model reverberating spiking neural network1 A. K. Vidybida Department of Synergetics, Bogolyubov Institute for Theoretical Physics Metrologichna str., 14-B, 03680 Kyiv, Ukraine E-mail: [email protected] Abstract Information about external world is delivered to the brain in the form of structured in time spike trains. During further processing in higher areas, information is subjected to a certain condensation process, which results in formation of abstract conceptual images of external world, apparently, represented as certain uniform spiking activity partially independent on the input spike trains details. Possible physical mechanism of condensation at the level of individual neuron was discussed recently. In a reverberating spiking neural network, due to this mechanism the dynamics should settle down to the same uniform/periodic activity in response to a set of various inputs. Since the same periodic activity may correspond to different input spike trains, we interpret this as possible candidate for information condensation mechanism in a network. Our purpose is to test this possibility in a network model consisting of five fully connected neurons, particularly, the influence of geometric size of the network, on its ability to condense information. Dynamics of 20 spiking neural networks of different geometric sizes are modelled by means of computer simulation. Each network was propelled into reverberating dynamics by applying various initial input spike trains. We run the dynamics until it becomes periodic. The Shannon's formula is used to calculate the amount of information in any input spike train and in any periodic state found. As a result, we obtain explicit estimate of the degree of information condensation in the networks, and conclude that it depends strongly on the net's geometric size. 1 Introduction The ability of a biological object to obtain enough information about external world is essential for the object's survival. The information is deliv- ered to the central nervous system through vari- ous sensory pathways in the form of spike trains. The pathways' information throughput has been studyed for a long time in theoretical [22] and ex- perimental [25] research. The paradigm of those research is consistent with the assumption that the best situation is when the brain receives maximum information from sensory systems. But compare this with [10]. Another paradigm, which is as well mature, [23], concentrates on self-organization of spike trains when primary sensory activity spreads to higher brain areas. Self-organization is ac- companied with information loss, [15], Indeed, a kind of standartization of activity evoked by var- ious primary sensory inputs is obsereved exper- imentally in higher brain areas during olfactory [6, 38] and auditory [4] perception. In visual sys- 1Accepted in the International Journal of Neural Systems, http://www.worldscinet.com/ijns/ijns.shtml 2 A. Vidybida tem, simple examples are the transformation of scene representation to viewpoint-invariant, [7] or retinotopic-invariant [32] coordinates. Here, spike trains in the optic nerve must depend on certain information (retinal position, viewpoint), which is removed in higher areas of conceptual repre- sentation. In the context of cognitive physiol- ogy, the process of reduction of information aimed at conceptual representation/recognition of exter- nal objects is known as information condensation, [21]. Usually pattern recognition phenomenon, which is closely related to information condensa- tion, is considered in parallel with training, see [12, 17, 24, 28, 30, 31, 39], learning, [1, 11, 13, 21], or other plasticity, [16], in the corresponding net- work. In a biological network, the learning mecha- nism involves biosynthesis [27] and is therefore very slow process, which requires seconds or minutes, [3]. At the same time, recognition of objects in vi- sual scene can be accomplished within 150 ms, or faster, citeThorpe1,Thorpe. During this short pe- riod of time, the network has constant structure, but it is the spiking activity which evolves within it from information-rich at the sensory periphery into information-poor, representing concrete enti- ties/concepts at higher brain areas. We now put a question: What could be the physical mechanism of information condensation in a spiking neural system? One possible mechanism, [37] which operates at the level of single neuron, was discussed, see Fig. 1, 2 and n. 2.1.12.1.1, be- low. In a reverberating spiking network, due to this mechanism the spiking dynamics should set- tle down to the same definite periodic activity in response to any stimulus from a definite set of var- ious inputs, and to another periodic activity in re- sponse to members of another set of inputs. If so, then the definite periodic dynamical state can be considered as an abstract representation of a fea- ture, which all stimuli from the definite set have in common. And this is just what is expected from the condensation of information. Our purpose in this work is to study how the ability of a simple spiking neural net to condense information in the above described sense depends on its physical parameter -- the net's geomet- ric size. For this purpose we simulate dynamics of a net composed of five binding neurons placed equidistantly on a circle. The circle's radius, R, characterizes the net's geometric size. The net is fully connected, and propagation velocity is taken the same for all connections and all values of R. Thus, the variations in R are expressed exclusively in the variations in the interneuronal propagation times. Initially, the net is stimulated by a spike train of input impulses triggering each of five neu- rons at times {t0, t1, . . . , t4}. Afterwards, the spik- ing dynamics is allowed to go freely until it settles down to a periodic one. This allows to figure out sets of input stimuli bringing about the same pe- riodic dynamics. The number of different periodic dynamics and the number of input stimuli in a set, which corresponds to a definite periodic dynamics, characterize the net's ability to condense informa- tion. We found that this ability depends strongly on the net's size. 2 Methods 2.1 The Binding Neuron Model The understanding of mechanisms of higher brain functions expects a continuous reduction from higher activities to lower ones, eventually, to ac- tivities in individual neurons, expressed in terms of membrane potentials and ionic currents. While this approach is correct scientifically and desirable for applications, the complete range of the reduc- tion is unavailable to a single researcher/engineer due to human brain limited capacity. In this con- nection, it would be helpful to abstract from the rules by which a neuron changes its membrane po- tentials to rules by which the input impulse sig- nals are processed into its output impulses. The coincidence detector, and temporal integrator are the examples of such an abstraction, see discus- sion by Konig et al., [20]. One more abstraction, the binding neuron (BN) model, is proposed as sig- nal processing unit, [34] which can operate as ei- ther coincidence detector, or temporal integrator, depending on quantitative characteristics of stim- ulation applied. This conforms with behavior of real neurons, see, e.g. work by Rudolph & Des- texhe, [26]. The BN model describes functioning of a neuron in terms of discret events, which are input and output impulses, and degree of tempo- ral coherence between the input events, see Fig. 1. Mathematically, this can be realized as follows. Each input impulse is stored in the BN for a fixed time, τ . The τ is similar to the tolerance inter- val discussed by MacKay, [23]. All input lines are excitatory and deliver identical impulses. The neu- ron fires an output impulse if the number of stored impulses, Σ, is equal or higher than the threshold value, N0. In this model, inhibition is expressed in decreased τ value. It is clear, that BN is triggered when a bunch of input impulses is received in a narrow temporal interval. In this case, the bunch could be considered as compound event, and the output impulse -- as an abstract representation of this compound event. One could treat this mech- anism as binding of individual input events into a single output event, provided the input events are coherent in time. Such interpretation is sug- Testing of Information Condensation 3 elementary event elementary event ... ... elementary event ✲ ✲ ✲ Binding of the elementary events based on their temporal coherence ④ elementary event for secondary neurons ✲ (represents the bound event) ✚✚❃ ✲✚ ❍ ❍❍❥ inhibition controls binding Figure 1: Signal processing in the binding neuron model.[35, 36] gested by binding of features/events in largescale neuronal circuits, [5, 8, 9, 36]. In our case, It would be interesting to characterize the BN input-output relations in the form of transfer func- tion, which allows exact calculation of output in terms of input. input is the se- quence of discrete arriving moments of standard impulses: Tin = {l1, l2, l3, l4, . . . }. The output is the sequence of discrete firing moments of BN: Tout = {f1, f2, . . . }. It is clear that Tout ⊂ Tin. The transfer function in our case could be the function σ(l), l ∈ Tin, which equals 1 if l is the firing mo- ment, l ∈ Tout, and 0 otherwise. For BN with threshold N0 the required function can be con- structed as follows. It is clear that the first N0 − 1 input impulses are unable to trigger neuron, there- fore σ(l1) = 0, . . . , σ(lN0−1) = 0. The next input is able to trigger if and only if all N0 inputs are coherent in time: σ(lN0 ) = 1 if and only if lN0 − l1 ≤ τ. In order to determine σ(lN0+k), k ≥ 1, one must take into acount all previous input moments, there- fore we use notation σTin instead of σ. The values of σTin (lN0 +k) can be determined recursively: σTin (lN0+k) = 1 if and only if lN0+k − lk+1 ≤ τ and σTin (li) = 0 for all i ∈ {k + 1, . . . , N0 + k − 1}. The function σTin describes completely the BN model for arbitrary threshold value N0 ≥ 2 . 2.1.1 Information Condensation in a Single Neuron It is worth noticing, that any firing (triggering) moment of a spiking neuron is determined by the moment of last input impulse, which just ensures that the triggering condition is satisfied. In a neu- ron, which needs more than one input impulse to fire, variations of temporal position of impulses, re- ceived just before the triggering one, do not influ- ence the moment of emitting the output impulse, provided those variations are in resonable limits and arrival moment of the triggering impulse re- mains the same. τ t25 t24 t23 t22 t21 t15 t14 t13 t12 t11 Figure 2: Example of two different inputs into a single neuron, which produce identical out- puts. Thus, different input spike trains can produce ex- actly the same output. This looks like if some de- tailes of the input stimulus, which is composed of several impulses, were reduced/condensed in the output. In the BN, the triggering condition is that the number of impulses in the BN's inter- nal memory equals to N0. Consider BN with N0 = 4, which is stimulated with two different spike trains with input impulses arrival times S1 = {t11, t12, t13, t14, t15}, t11 < t12 < t13 < t14 < t15, and S2 = {t21, t22, t23, t24, t25}, t21 < t22 < t23 < t24 < t25. Let the arrival moments satisfy the fol- lowing conditions: t14 − t11 < τ, t25 − t22 < τ, t24 − t21 > τ, t14 = t25 = to . 4 A. Vidybida In this case, both S1 and S2, if fed to the BN, will produce exactly the same output, namely, the sin- gle impulse at moment to. This is illustrated in Fig. 2. 2.2 The Network Model As a reverberating spiking neural net we take the net of five neurons placed equidistantly at a circle of radius R, see Fig, 3. Each neuron has threshold N0 = 4, and internal memory, τ = 10 ms. The net is fully connected. The connection lines are characterized with length and propagation veloc- ity, v, which is the same in all lines. For R fixed, there are two types of connection line, the short one, with propagation delay d, and the long one with propagation delay D. Each neuron has addi- tionally the external stimulus input line, which is used to start the net dynamics. Single impulse in the stimulus input line delivers to its target neuron just threshold excitation. This causes firing at the moment of the stimulus impulse arrival. For nu- merical simulations we use 20 networks of different sizes, see Table 1. The propagation velocity in any interconnection line is taken v = 0.1 m/s. 2.2.1 Numerical Simulation As programming language we use Python under Linux operating system. The dynamics was mod- elled by advancing time with step dt = 200 µs. The delay values d and D, when measured in the dt units, were rounded to the nearest from below integers, see Table 1. As a result, the simulating program operates in whole numbers with no round- ing errors involved. Each single tick of the program, the network's tick, advances time by dt, and consists of three par- tial ticks, which are performed in the given order. Namely, (i) input tick, which advances time in the input lines, (ii) axonal tick, which advances time in the internal connection lines, (iii) neuronal tick, which advances time in the neurons. This manner of updating states can be treated as synchronous in a sense that each component of the network has the same physical time when the network's tick is complete. On the other hand, viewed as interneuronal communication process, the dynamics should be treated as asynchronous due to nonzero propagation delays. During the step (ii), a neuron can get impulse into its internal memory. If a neuron appears in the state "Fire" as a result of the network tick, then the output impulse it produces can appear in the connection lines only during the next network tick. This introduces effective delay of one dt between delivering the triggering impulse to a neuron and emitting output impulse by that neuron. 2.3 Data Acquisition Algorithm 2.3.1 Set of Stimuli The net was entrained to reverberating dynamics by applying initial input spike train of five im- pulses, one triggering impulse per neuron, at times (in dt units) {t0 = 1, t1, t2, t3, t4}. The triggering moment of neuron # 0 is taken 1 for all stimuli in order to exclude rotational symmetry between the stimuli applied. Other four triggering moments run independently through the set {1, 2, . . . , tmax}, where tmax is choosen proportional to R for each net size. In choosing tmax, we follow two different paradigms. In the first paradigm of short stim- uli we restrict the overall duration of the stim- ulus train with the value tmax = d. Thus, the set of stimuli has d4 different stimuli. If ti ≤ d, i = 0, . . . , 4, then any neuron in the net never obtains impulse from other neurons before it ob- tains its external input stimulation. In the second paradigm of extended stimuli, we restrict the over- all duration of the stimulus train with the value tmax = M , which is about three times longer than d for each network (see Table 1). Here, all stim- uli {t0 = 1, t1, t2, t3, t4}, which were presented to a network, cover the set of M 4 different trains, which equals from 625 different stimuli for net #1 to 100 000 000 different stimuli for net #20 (see Table 1). The stimuli were sampled in accordance with standard algorithm of 4-digit counter. Namely, we started from stimulus {1, 1, 1, 1, 1}, the next stim- ulus is obtained by advancing t1 by 1, and so on. The stimulus next to {1, M, 1, 1, 1} is {1, 1, 2, 1, 1}, the one next to {1, M, M, 1, 1} is {1, 1, 1, 2, 1} and so on, until stimulus {1, M, M, M, M } is presented. In the extended paradigm, the late external in- put impulse can enter corresponding neuron after it received impulses from neurons already triggered by early external input impulses. The second paradigme is in concordance with visual information processing, [2] where activity from higher brain areas, which was invoked due to visual stimulation at earlier time, is retroinjected to areas V1 and V2 in the primary visual cortex, where it interacts with activity invoked by visual input at later time during perception. 2.3.2 Figuring out Periodic States After the last input impulse from the train {t0 = 1, t1, t2, t3, t4} reachs its target neuron, the pro- gram begins appending at each time step the in- stantaneous state of the net to a Python list. The instantaneous state consists of states of all 20 con- nection lines and all 5 neurons (see Fig. 4). Before appending, the program checks if the current in- stantaneous state was already included in the list. Testing of Information Condensation 5 t 2 d n2 t1 n1 D n3 t 3 n4 t 4 t0 = 1 n0 Figure 3: The network, used for simulations. Here {t0, t1, t2, t3, t4} -- is the input spike train, d, D -- are the propagation delays in the connection lines. Any line can be either empty, or propagating one impulse. net # R, mm 0.029 1 d, dt D, dt M , dt net # R, mm 0.314 1 2 5 11 d, dt D, dt M , dt 18 29 55 2 3 4 5 6 7 8 9 10 0.057 0.086 0.114 0.143 0.171 0.200 0.229 0.257 0.286 3 5 10 12 5 8 15 13 6 10 20 14 8 13 25 15 10 16 30 16 11 19 35 17 13 21 40 18 15 24 45 19 16 27 50 20 0.343 0.371 0.400 0.429 0.457 0.486 0.514 0.543 0.571 20 32 60 21 35 65 23 38 70 25 40 75 26 43 80 28 46 85 30 48 90 31 51 95 33 54 100 Table 1: Dimensions of networks used for simulations; dt = 200 µs. net # 1 31 2 61 104 123 net # 11 301 494 12 331 544 3 91 154 187 246 13 361 584 4 111 184 227 296 14 391 634 5 141 234 287 376 15 411 6 171 284 347 456 16 441 7 201 324 407 526 17 471 8 221 364 447 9 251 414 507 10 281 454 567 18 491 19 521 20 551 Table 2: Distinct periods in dt units of found periodic states in the short stimuli paradigm. Superscript denotes the number of different periodic states with this period. 6 A. Vidybida If it was, then the periodic dynamical state is found with its complete cyclic trajectory covered by in- stantaneous states between the inclusion and the last record in the list, inclusively. Measures are taken in order not to count the same cyclic trajec- tory, which was entrained at its different points, as different periodic states. The data for each net were stored in two MySQL tables. Table STATES included the se- rial number of any periodic state found, one ele- ment from the corresponding cyclic trajectory, and period of the state (see Fig. 4). Single record in the INPUTS table included the input stimulus {t0 = 1, t1, t2, t3, t4}, the serial number of the pe- riodic state it leads to (this number is 0 for fading dynamics), and relaxation time, namely, the time, which is spent between the last external input im- pulse is delivered and the net's entrance moment into the periodic regime. 3 Results 3.1 Characterization of Periodic States Found After initial stimulation, networks from #1 to #7 entrain to periodic activity after any stimulation, and networks #8 to #20 either entrain to peri- odic dynamics, or stops from any activity after some time. This takes place for both short and extended stimuli.The number of different periodic states found for each network is shown in Fig. 5. Here, the maximal number of periodic states ob- tained with short stimuli is 18, and this number is achieved for net numbers from 3 to 7. Exact values of periods, and number of different periodic states with this period is shown in Table 2 for short stim- uli. We omit similar table for extended stimuli. The maximal number of different periodic states obtained with extended stimuli is 485, which is achieved in the net number 9. The maximal num- ber of different periodic states with the same pe- riod is 294 for period 50·dt in net #9. It should be mentioned that two periodic states, which can be turned into eachother by suitable renumeration of neurons, were considered as different. It is evident that in the network of five neurons with threshold 4, each neuron is triggered the same number of times during period. Indeed, expect that neuron n4 fires k4 times, and any of the other four fires less during period: k4 > ki, i = 0, 1, 2, 3. In order to be triggered k4 times, n4 must obtain not less than 4k4 input impulses during period. But it can obtain only k0 +k1 +k2 +k3, which is less than required. Similarly, situation when k4 = k3 and k3 > ki, i = 0, 1, 2 leads to contradiction, and so on. This number of triggering is either 1 or 2 for trajectories found, see examples in Fig. 6. Some nets have only one periodic state, which coresponds to synchronous firing of all 5 neurons and symmet- rical states of connection lines at any moment of time. This is the case for nets number 1 and from 15 to 20 for short stimuli, and for nets number 19 and 20 for extended stimuli. 3.2 Condensation of Information In order to estimate the degree of information con- densation in the course of transformation of an ex- ternal spike train into a certain periodic state of the net, one needs to calculate information amount, which is delivered by specifying a spike train, and which is delivered by specifying the state, it leads to. 23 0, 1 4 2, 3 0, 4 state #319 50 1 state #107 50 0, 3 1, 2 4 0, 3 1, 2 4 state #199 66 0 0 0 Figure 6: Examples of periodic states found for net number 9 in the extended stimuli paradigm. Spikes indicate the firing moments, labels near each spike give numbers of neurons, firing at this moment. The two upper trains show states with period 10 ms, the lower one - with period 13.2 ms. This can be done by wellknown Shannon's formula [29] H = − X pi log2 pi, i (1) where pi is the probability to obtain case num- ber i from a set of cases. At the input end we have the set of d4, or M 4 different external input spike trains. In our statement of the problem, it is natural to consider all external input trains as equally probable. If so, then information delivered by specifying certain train is Hs = 4 log2 d, for the short stimuli paradigm, and He = 4 log2 M, (2) (3) for the extended stimuli paradigm. While estimating information, delivered by specifying certain periodic state, one should take into account that probabilities of different periodic Testing of Information Condensation 7 mysql> select * from STATES_5_9 where num=269; +------+------------------------------------------------------------------------------ ----------------------------------------------+--------+ num state period +------+------------------------------------------------------------------------------ ----------------------------------------------+--------+ 269 0 1 1 0 8 8 17 17 24 15 15 24 8 8 0 0 8 17 17 8 0 False False 0 False False 0 False False 0 False False 0 False True 49 1 82 +------+------------------------------------------------------------------------------ ----------------------------------------------+--------+ 1 row in set (0.02 sec) mysql> Figure 4: Example of single record in the MySQL table STATES. The first field (num) is numerical, and gives the serial number of periodic state found. The second field (state) is a string, which describes instantaneous state from the periodic state found (a point from the cyclic trajectory, which represents the whole trajectory, or periodic state). The first 20 numbers in the string describe states of all connection lines: '0' means that the line is empty, positive number specifies after how many ticks the propagating impulse will rich the targeted neuron. The next five chunks confined between the "next line" symbols describe states of neurons. The first number in each chunk is the "kick" -- the total number of impulses obtained by neuron after the axonal tick was complete. During the neuronal tick, corresponding to that axonal tick, the "kick" is utilized and set to zero. The next boolean in the chunk indicates if the neuron is in the "Fire" state. The next boolean indicates if the neuron has any impulses in its internal memory. If it has, then next couples of numbers (up to three couples) describe those impulses. In this example, neuron #4 has 1 impulse with time to leave 49·dt. The third field (period) specifies period (in dt units) of this periodic state. s e t a t s c i d o i r e p f o r e b m u n 20 15 10 5 0 2 4 6 8 10 net # 12 14 16 18 20 s e t a t s c i d o i r e p f o r e b m u n 500 450 400 350 300 250 200 150 100 50 0 2 4 6 8 10 net # 12 14 16 18 20 Figure 5: Number of different periodic states found for each net with short (left panel), and extended (right panel) stimuli. 8 A. Vidybida states are not the same. In order to calculate prob- ability pn of a periodic state Cn, we calculate the number of input spike trains leading to the Cn, namely, Tn, and divide this number by the total amount of different input stimuli (see histogram for Tn in Fig. 7) pn = Tn d4 , for the short stimuli paradigm, and pn = Tn M 4 , (4) (5) for the extended stimuli paradigm. Then we use Eq. (1) with probabilities of individual periodic states found in accordance with (4), (5) to cal- culate information which should be ascribed to any periodic state. In this calculations, we treat uniformly with others the external input stimuli, which lead to fading dynamics. Correspondingly, the state with no activity is treated uniformly with periodic states. This is in the contrast with data presented in Fig. 5, and Table 2, where the state with no activity is excluded. n i b e n o n i s n i a m o d f o r e b m u n 90 80 70 60 50 40 30 20 10 0 s e t a t s c i d o i r e p f o n o i t a m r o f n i 8 7 6 5 4 3 2 1 0 e s 2 4 6 8 10 net # 12 14 16 18 20 Figure 8: Dependence of information amount, which is ascribed to periodic state, on the net size. Curve 'e' corresponds to extended stimuli paradigm, 's' -- to the short stimuli one. information of a periodic state, calculated in accor- dance with Eq. (1) with probabilities found due to Eqs. (4), (5), varies between 6.93 and 7.33 bits for extended stimuli, and between 3.17 and 3.46 for short stimuli. In the plateau, the degree of in- formation condensation calculated as input infor- mation divided by the periodic state information, varies between 9.02 and 12.27 for extended stim- uli, and between 11.2 and 19.31 for short stimuli. Out of a plateau range, for greater net sizes the amount of information in a periodic state drops sharply due to simplification of the set of peri- odic states. Namely, the total number of periodic states decreases to one, with the second one with no activity, and the probability is distributed very unevenly between this two states. This leads to the degree of information condensation as high as 41000 for extended stimuli, and 690 for short ones. 0 10000 20000 30000 40000 size of conceptual domain 4 Discussion Figure 7: Histogram of conceptual domain sizes for net #9 in extended stimuli paradigm. The bin size is 518. 23 domains with sizes from 50764 to 193732 are not presented. As it could be expected, the information amount in an external input spike train increases as logarithm of the net size, in accordance with Eqs. (2), (3), varying from 25 to 81 bits for short paradigm, and from 37 to 106 bits for extended paradigm. Information, which could be ascribed to a periodic state, depends on the net size in a more complicated manner, see Fig. 8. A remark- able feature is a kind of plateau between net #3 and #9 for both short, and extended stimulation paradigme. In the plateau, the Any reverberating spiking neural net can repre- sent complicated dynamical behavior. If the net's instantaneous states can be described with whole numbers, then the net will inevitably either en- train to periodic dynamics, or stop its activity at all. In this study, it is appeared that a very simple net of Fig. 3 can be engaged into a considerably large set of different periodic activities. It is not clear which part of all possible in this net periodic states was discovered in our simulation. As it fol- lows from comparison between short and extended stimuli paradigm, the number of different periodic states found increases with increasing range of in- put stimuli. Certainly, this increase must saturate somewhere. This is because any two different peri- odic regimes are represented by their cyclic trajec- tories, which has no common points (instantaneous states). On the other hand, the total number of instantaneous states the network can have is finite Testing of Information Condensation 9 due to finiteness of the set of states of each element the network is composed of. The number of periodic states found in a net depends on the net's geometric size. Variations in the net's size display themselves exclusively in variations of interneuronal propagation delays d and D. On the other hand, the duration of neu- ronal internal memory, τ , is the same for net of any size. Thus, it is namely the relationships between the times an impulse spends for travelling between neurons, and time it is allowed to spend in a neu- ron waiting for additional impulses, which controls possible number of periodic states. It is worth noticing that each net has one com- pletely synchronized periodic state. The com- pletely synchronozed state is stable and achieved during finite time. This is in the contrast to the case of pulse-coupled oscillators with delayed exci- tatory coupling, (see, e.g. Ref. [40]). In the four-dimensional set of stimuli we used, the neighbouring stimuli differ from eachother by one dt in one of four dimensions. This can be treated as analogous representation of some reality. The set of periodic states should be considered as a set of discrete entities due to qualitative differ- ence between any two states. This conforms with a paradigm discussed in cognitive physiology, [21]. The process of transformation of initial analogous inputs into a discret set of periodic states implies a loss of information and can be treated as conden- sation of information. If we take a set of input stimuli, any of which leads to the same periodic state, then that periodic state can be considered as an abstract/conceptual representation of a feature, which all stimuli from the set have in common, and the corresponding set could be named as "conceptual domain". What kind of feature or concept does the conceptual do- main represent? If our net was trained to recognize a certain real feature, then it would be that feature. In the context of this study, the common feature is that all stimuli from the conceptual domain engage namely this net into namely this periodic dynam- ics. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 42 0 349 2 8 125 42 262 273 42 0 349 300 8 125 42 262 273 41 256 42 8 172 182 371 42 125 0 133 262 273 270 201 41 40 92 35 133 270 178 178 269 268 267 266 265 264 221 220 221 264 265 266 267 268 269 270 35 270 270 203 212 202 92 259 201 40 259 260 259 92 256 42 125 42 8 172 182 371 41 133 262 273 356 354 352 350 92 41 256 348 133 270 178 35 35 270 178 269 268 267 266 265 264 221 220 221 264 265 266 267 268 269 270 256 178 270 228 45 132 256 177 178 170 270 228 270 178 181 186 190 193 276 275 274 240 239 Figure 9: 2-dimensional cross-section of the 4- dimensional space of inputs for network #9 in 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 358 295 2 295 330 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 300 302 301 297 301 295 295 0 extended stimuli paradigm. left, cross-section is made with plane (t3 = 23, t4 = 23), right, (t3 = 21, t4 = 23). Origin for (t1, t2) is in the upper left corner. Both t1 and t2 run through the set {1, 2, . . . , 45} of values. Num- bers in polygons indicate serial numbers of cor- responding periodic states. It would be interesting to have a look on the topology of the conceptual domains in the 4- dimensional space of all stimuli. For this pur- pose we figured out 2-dimensional cross-section of the input stimuli (see Fig. 9). In the cross- section, a typical conceptual domain is composed of several coherent clusters disconnected with ea- chother. The histogram of sizes of conceptual do- mains found is given in Fig. 7. Decomposition of the whole set of input stimuli into a number of conceptual domains represented by corresponding periodic dynamics resembles ap- proach in analyses of multivariate datasets, see re- view in [14]. The difference is that here the 4- dimensional dataset (a conceptual domain) is rep- resented by unidimensional cyclic trajectory, which corresponds to the domain, and the trajectory is composed of points/vectors, which have other di- mension than the dataset vectors (see Fig. 4). Nevertheless, having the network, the whole cyclic trajectory can be reproduced starting from its any single point. Thus, here the datasets are reduced down to individual points. This is in concordance with the information condensation idea. Why do we stick ourselves with namely the pe- riodic dynamics? The answer is related to the memory/learning problem, even if we do not con- sider any plasticity in this study. It is known [3] that modification of synaptic strength may hap- pen due to repetitive delivery of impulses to those synapses. Periodic dynamical states are just well suited for such repetitive delivery. All other dy- namical behaviors are of transient type, and have less chances to cause plastic changes in biological network. On the other hand, successful perception expects ability to report about what was perceived, which is impossible without memory. Unfortunately, we cannot draw this biological analogy too far. Real biological network includes not only excitatory, but about 4% of inhibitory neurons, [18]. Representing this characteristic in a model network requires to have at least 25 neu- rons in it. This dramatically increases the number of possible external stimuli, which requires a quali- tative change in our approach as regards the speed of simulations and analysis of obtained data. Finally, what happens if we use another neu- ronal model in the network? Our opinion is that re- sults will be qualitatively similar. Using the bind- ing neuron here is natural, since it represents in 10 A. Vidybida refined form what a spiking neuron does with sig- nals it receives. Additionally, the BN model easily allows to develop a program operating in whole numbers. This excludes possible dynamical arte- facts due to rounding errors. In future work, it would be interesting to com- pare results, if another spiking neuron model is used in the network, to study the topology of conceptual domains and how the topology could change if a plasticity is introduced in the network model. 5 Conclusions A network composed of spiking neurons is able to condense information due to the fact that differ- ent initial stimuli could lead the network to the same periodic dynamics. This happens by means of initial/basic condensation of information in spik- ing neurons, as it is described in n. 2.1.12.1.1, above. The network's geometric size, which deter- mines the interneuronal transmission delays, has considerable influence on the net's ability to con- dense information, mainly due to influence on the number of different periodic states the network can have, see Fig. 5. The latter has influence on the amount of information, which should be ascribed to a single periodic state, see Fig. 8. As a re- sult, the degree of information condensation varies between 9 and 41 000, see n. 3.23.2 for details. The networks considered here are too primi- tive to have reliable biological implications. At the same time, numerical parameters, see Table 1 and n. 2.22.2, such as network sizes and spike propaga- tion velocity, are taken corresponding to biological data. The threshold value 4 does not contradict to biological reality, as experimentally registered thresholds are between 1 and 300. The BN inter- nal memory duration, τ , is commensurable with halfdecay time of the excitatory postsynaptic po- tentials (EPSP). Thus, in the framework of this extremely simple model, one could expect that the ability of biological neural network to condense in- formation should depend on its geometric size, or on the relationships between interneuronal trans- mission delays and the EPSP halfdecay time. Acknowledgments This work was supported by the Program of basic research of the National Academy of Science of Ukraine. Content of this work was partially published in an abstract form in the abstract book of the 2nd International Biophysics Congress and Biotechnology at GAP & 21th National Biophysics Congress, (5-9 Oct. 2009) Diyarbakır, Turkey, http://www.ibc2009.org/ References [1] Acharya, R., Chua, E.C.P., Chua, K.C., Min, L.C., and Tamura, T. (2010), Anal- ysis and Automatic Identification of Sleep Stages using Higher Order Spectra, Interna- tional Journal of Neural Systems, 20:6. [2] J. Bullier, "Integrated model of visual pro- cessing," Brain Res. Rev., 36, 96 -- 107 (2001). [3] D. V. Buonomano and M. M. Merzenich, "Cortical plasticity: to maps," Annu. Rev. Neurosci., 21, 149 -- 186 (1998). from synapses [4] P. Cariani, "Temporal codes, timing nets, and music perception," J. New Music Res., 30, 107 -- 135 (2001). [5] A. R. Damasio, "The brain binds entities and events by multiregional activation from con- vergence zones," Neural Comput., 1, 123 -- 132 (1989). [6] P. Duchamp-Viret and A. Duchamp, "Odor processing in the frog olfactory system," Prog. Neurobiol., 53, 561 -- 602 (1997). [7] R. Durbin and G. Mitchison, "A dimension reduction framework for understanding cor- tical maps," Nature, 343, 644 -- 647 (1990). [8] R. Eckhorn, R. Bauer, W. Jordan, M. Brosch, W. Kruse, M. Munk and H. J. Reit- boeck, "Coherent oscillations: a mechanism for feature linking in the visual cortex?," Biol. Cybern., 60, 121 -- 130 (1988). [9] A. K. Engel, P. Konig, A. K. Kreiter, C. M. Gray and W. Singer, "Temporal coding by coherent oscillations as a potential so- lution to the binding problem: physiolog- ical evidence," In Schuster, H.G., Singer, W. (ed), Nonlinear Dynamics and Neuronal Networks, 3 -- 25. VCH Weinheim (1991). [10] J. Feldman, "Ecological expected utility and the mythical neural code," Cognitive Neuro- dynamics, 4, 25 -- 35 (2010). [11] Ghosh-Dastidar, S. and Adeli, H. (2007), Im- proved Spiking Neural Networks for EEG Classification and Epilepsy and Seizure De- tection, Integrated Computer-Aided Engi- neering, Vol. 14, No. 3, pp. 187-212. [12] S. Ghosh-Dastidar and H. Adeli, "Spiking neural networks," Intern. J. Neural Sys. 19, 295 -- 308 (2009). [13] S. Ghosh-Dastidar and H. Adeli, A New Supervised Learning Algorithm for Multiple Spiking Neural Networks with Application in Epilepsy and Seizure Detection, Neural Net- works 22, 1419 -- 1431 (2009). Testing of Information Condensation 11 [14] A. Gorban and A. Zinovyev, "Principal man- ifolds and graphs in practice: from molecu- lar biology to dynamical systems," Intern. J. Neural Sys., 20, 219 -- 232 (2010). [15] H. Haken, "Brain Dynamics. Synchroniza- tion and Activity Patterns in Pulse-Coupled Neural Nets with Delays and Noise," Springer, Berlin 2007. [16] J. Iglesias and A. E. P. Villa, "Emergence of preferred firing sequences in large spik- ing neural networks during simulated neu- ronal development," Intern. J. Neural Sys., 18, 267 -- 277 (2008). [17] Johnston, S.P., Prasad, G.,Maguire, L. and McGinnity, T.M. (2010), An FPGA Hard- ware/software co-design methodology - to- wards evolvable spiking networks for robotics application, Intern. J. Neural Sys., 20:6. [18] E. R. Kandel, J. H. Schwartz and T. M. Jes- sell, "Principles of Neural Science," (Fourth ed.), New York: McGraw-Hill (2000). [19] H. Kirchner and S. J. Thorpe, "Ultra-rapid object detection with saccadic eye move- ments: visual processing speed revisited," Vision Res., 46, 1762 -- 1776 (2006). [20] P. Konig, A. K. Engel and W. Singer, "Inte- grator or coincidence detector? The role of the cortical neuron revisited," Trends Neu- rosci., 19, 130 -- 137 (1996). [21] P. Konig and N. Kruger, "Symbols as self- emergent entities in an optimization process of feature extraction and predictions," Biol. Cybern., 94(4), 325 -- 334 (2006). [22] D. M. MacKay and W. S. McCulloch, "The limiting information capacity of a neuronal link," Bull. Math. Biophys., 14, 127 -- 135, (1952). [23] D. M. MacKay, "Self-organization in the time domain," In M. C. Yovitts, G. T. Jacobi and G. D. Goldstein (ed), Self-Organizing Systems, Spartan Books, Washington, 37 -- 48 (1962). [24] Nichols, E., McDaid, L.J., and Siddique, N.H. (2010), Case Study on Self-organizing Spiking Neural Networks for Robot Naviga- tion, Intern. J. Neural Sys., 20:6. [25] Ch. L. Passaglia and J. B. Troy, "Informa- tion transmission rates of cat retinal gan- glion cells," J. Neurophysiol., 91, 1217 -- 1229 (2004). [26] M. Rudolph and A. Destexhe, "Tuning neo- cortical pyramidal neurons between integra- tors and coincidence detectors, " J. Comput. Neurosci., 14, 239 -- 251 (2003). [27] M. T. Scharf, N. H. Woo, K. M. Lattal, Z. Young, P. V. Nguyen and T. Abel, "Protein synthesis is required for the enhancement of long-term potentiation and long-term mem- ory by spaced training," J. Neurophysiol., 87 2770 -- 2777 (2002). [28] Schliebs, S., Kasabv, N., and Defoin-Platel, M. (2010), On the Probabilistic Optimiza- tion of Spiking Neural Networks, Intern. J. Neural Sys., 20:6. [29] C. E. Shannon, "A mathematical theory of communication," Bell System Technical J., 27, 379 -- 423 and 623 -- 656, July and October, (1948). [30] Soltic and S. Kasabov, N. (2010), Knowledge extraction from evolving spiking neural net- works with rank order population coding , Intern. J. Neural Sys., 20:6. [31] Strain, T.J., McDaid, L.J., Maguire, L.P., and T.M. McGinnity, T.M. (2010), An STDP Training Algorithm for a Spiking Neu- ral Network with Dynamic Threshold Neu- rons, Intern. J. Neural Sys., 20:6. [32] S. Tang, R. Wolf, S. Xu and M. Heisenberg, "Visual pattern recognition in Drosophila is invariant for retinal position," Science, 305, 1020 -- 1022 (2004). [33] S. Thorpe, D. Fize and C. Marlot, "Speed of processing in the human visual system," Nature, 381, 520 -- 522 (1996). [34] A. K. Vidybida, "Neuron as time coherence discriminator," Biol. Cybern., 74, 539 -- 544 (1996). [35] A. K. Vidybida, "Information process- ing in a pyramidal-type neuron," Proc. BioNet'96 Biologieorientierte Informatik und pulspropagierende Netze, 3-d Workshop. Ed.: Heinz G., pp. 96 -- 99. (1996). [36] A. K. Vidybida, "Inhibition as binding con- troller at the single neuron level," BioSys- tems, 48, 263 -- 267 (1998). [37] A. K. Vidybida, "Output stream of binding neuron with instantaneous feedback," Eur. Phys. J. B, 65, 577 -- 584 (2008); Eur. Phys. J. B, 69, 313 (2009). [38] M. Wehr and G. Laurent, "Odor encoding by temporal sequences of firing in oscillat- ing neural assemblies" Nature, 384, 162 -- 166 (1996). [39] B. Widrow, "Generalization and information storage in networks of adaline 'neurons' " in: Yovitz,M.C., Jacobi,G.T., Goldstein,G.(ed), Self-Organizing Systems, 435 -- 461 (1962). 12 A. Vidybida [40] W. Wu and T. Chen, "Impossibility of asymptotic synchronization for pulse- coupled oscillators with delayed excitatory coupling," Intern. J. Neural Sys., 19, 425 -- 435 (2009). s n i a m o d f o r e b m u n 5 4 3 2 1 0 1 10 100 1000 10000 100000 1e+06 size of conceptual domain
0905.2935
3
0905
2010-01-18T02:39:52
Experimental demonstration of associative memory with memristive neural networks
[ "q-bio.NC", "cond-mat.mes-hall", "q-bio.MN" ]
When someone mentions the name of a known person we immediately recall her face and possibly many other traits. This is because we possess the so-called associative memory, that is the ability to correlate different memories to the same fact or event. Associative memory is such a fundamental and encompassing human ability (and not just human) that the network of neurons in our brain must perform it quite easily. The question is then whether electronic neural networks (electronic schemes that act somewhat similarly to human brains) can be built to perform this type of function. Although the field of neural networks has developed for many years, a key element, namely the synapses between adjacent neurons, has been lacking a satisfactory electronic representation. The reason for this is that a passive circuit element able to reproduce the synapse behaviour needs to remember its past dynamical history, store a continuous set of states, and be "plastic" according to the pre-synaptic and post-synaptic neuronal activity. Here we show that all this can be accomplished by a memory-resistor (memristor for short). In particular, by using simple and inexpensive off-the-shelf components we have built a memristor emulator which realizes all required synaptic properties. Most importantly, we have demonstrated experimentally the formation of associative memory in a simple neural network consisting of three electronic neurons connected by two memristor-emulator synapses. This experimental demonstration opens up new possibilities in the understanding of neural processes using memory devices, an important step forward to reproduce complex learning, adaptive and spontaneous behaviour with electronic neural networks.
q-bio.NC
q-bio
Experimental demonstration of associative memory with memristive neural networks Yuriy V. Pershin and Massimiliano Di Ventra 1 0 1 0 2 n a J 8 1 ] . C N o i b - q [ 3 v 5 3 9 2 . 5 0 9 0 : v i X r a Abstract—Synapses are essential elements for computation and information storage in both real and artificial neural systems. An artificial synapse needs to remember its past dynamical history, store a continuous set of states, and be ”plastic” according to the pre-synaptic and post-synaptic neuronal activity. Here we show that all this can be accomplished by a memory-resistor (memristor for short). In particular, by using simple and inexpensive off- the-shelf components we have built a memristor emulator which realizes all required synaptic properties. Most importantly, we have demonstrated experimentally the formation of associative memory in a simple neural network consisting of three electronic neurons connected by two memristor-emulator synapses. This experimental demonstration opens up new possibilities in the understanding of neural processes using memory devices, an important step forward to reproduce complex learning, adaptive and spontaneous behavior with electronic neural networks. Index Terms—Memory, Resistance, Neural network hardware, Neural networks. I. INTRODUCTION W HEN someone mentions the name of a known person we immediately recall her face and possibly many other traits. This is because we possess the so-called asso- ciative memory - the ability to correlate different memories to the same fact or event [1]. This fundamental property is not just limited to humans but it is shared by many species in the animal kingdom. Arguably the most famous example of this are experiments conducted on dogs by Pavlov [2] whereby salivation of the dog’s mouth is first set by the sight of food. Then, if the sight of food is accompanied by a sound (e.g., the tone of a bell) over a certain period of time, the dog learns to associate the sound to the food, and salivation can be triggered by the sound alone, without the intervention of vision. Since associative memory can be induced in animals and we, humans, use it extensively in our daily lives, the network of neurons in our brains must execute it very easily. It is then natural to think that such behavior can be reproduced in artificial neural networks as well - a first important step in obtaining artificial intelligence. The idea is indeed not novel and models of neural networks have been suggested over the years that could theoretically perform such function [3], [4], [5], [6]. However, their experimental realization, especially in the electronic domain, has remained somewhat difficult. The reason is that an electronic circuit that simulates a neural network capable of associative memory needs two important components: neurons and synapses, namely connec- tions between neurons. Ideally, both components should be of nanoscale dimensions and consume/dissipate little energy so that a scale-up of such circuit to the number density of a typical human brain (consisting of about 1010 synapses/cm2) could be feasible. While one could envision an electronic version of the first component relatively easily, an electronic synapse is not so straightforward to make. The reason is that the latter needs to be flexible (”plastic”) according to the type of signal it receives, its strength has to depend on the dynamical history of the system, and it needs to store a continuous set of values (analog element). In the past, several approaches with different levels of ab- straction were used in order to implement electronic analogues of synapses [7]. For instance, one of the first ideas involved the use of three-terminal electrochemical elements controlled by electroplating [8]. While some of these approaches do not involve synaptic plasticity at all [9], [10], [11], [12], the latter is generally implemented using a digital (or a combination of analog and digital) hardware [13], [14], [15], [16], [17], [18], [19]. The common feature of synaptic plasticity realizations is the involvement of many different circuit elements (such as transistors) and, therefore, occupation of a significant amount of space on a VLSI chip. Thus, the amount of electronic synapses in present VLSI implementations is much lower than the amount of synapses relevant to actual biological systems. Novel, radically different approaches to resolve this issue would be thus desirable. A recently demonstrated resistor with memory (memristor Input 1 “ i h d” “sight of food” f f N 1 1 Input 2 p “sound” N 2 S S 1 1 S S 2 N N 3 Output “salivation” Yu. V. Pershin is with the Department of Physics and Astronomy and USC Nanocenter, University of South Carolina, Columbia, SC, 29208 e-mail: [email protected]. M. Di Ventra is with the Department of Physics, University of California, San Diego, La Jolla, California 92093-0319 e-mail: [email protected]. Manuscript received November XX, 2009; revised January YY, 2010. Fig. 1. Artificial neural network for associative memory. Real neurons and their networks are very complex systems whose behavior is not yet fully understood. However, some simple brain functions can be elucidated studying significantly simplified structures. Here, we consider three neurons (N1, N2 and N3) coupled by two memristive synapses (S1 and S2). The output signal is determined by input signals and strengths of synaptic connections which can be modified when learning takes place. for short [20], [21]) based on TiO2 thin films [22], [23] offers a promising realization of a synapse whose size can be as small as 30×30×2 nm3. Using TiO2 memristors, a fabrication of neuromorphic chips with a synapse density close to that of the human brain may become possible. Memristors belong to the larger class of memory-circuit elements (which includes also memcapacitors and meminductors) [24], [25], namely circuit elements whose response depends on the whole dynamical history of the system. Memristors can be realized in many ways, ranging from oxide thin films [22], [23], [26] to spin memristive systems [27]. However, all these realizations are limited to the specific material or physical property respon- sible for memory, and as such they do not easily allow for tuning of the parameters necessary to implement the different functionalities of electronic neural networks. In the present paper, we describe a flexible platform al- lowing for simulation of different types of memristors, and experimentally show that a memristor could indeed function as a synapse. We have developed electronic versions of neurons and synapses whose behavior can be easily tuned to the func- tions found in biological neural cells. Of equal importance, the electronic neurons and synapses were fabricated using inexpensive off-the-shelf electronic components resulting in few dollars cost for each element, and therefore can be realized in any electronic laboratory. Clearly, we do not expect that with such elements one can scale up the resulting electronic neural networks to the actual brain density. However, due to their simplicity reasonably complex neural networks can be constructed from the two elemental blocks developed here and we thus expect several functionalities could be realized and studied. For the purpose of this paper we have built the neural network shown in Fig. 1. We have then shown that such circuit is capable of associative memory. In this network, two input neurons are connected with an output neuron by means of synapses. As an example of the functionality that this network can provide, we can think about the animal memory we have described above [2] in which the first input neuron (presumably located in the visual cortex) activates under a specific visual event, such as ”sight of food”, and the second input neuron (presumably located in the auditory cortex) activates under an external auditory event, such as a particular ”sound”. Depending on previous training, each of these events can trigger ”salivation” (firing of the third output neuron). If, at a certain moment of time, only the ”sight of food” leads to ”salivation,” and subsequently the circuit is subjected to both input events, then, after a sufficient number of simultaneous input events the circuit starts associating the ”sound” with the ”sight of food”, and eventually begins to ”salivate” upon the activation of the ”sound” only. This process of learning is a realization of the famous Hebbian rule stating, in a simplified form, that ”neurons that fire together, wire together”. II. RESULTS AND DISCUSSION A. Electronic neuron Biological neurons deal with two types of electrical signals: receptor (or synaptic) potentials and action potentials. A stimu- 2 a V IN+ ADC I I V IN- NC O O -2.5V +2.5V Microcontroller b ) V ( e g a t l o V 5 4 3 2 1 0 -1 -2 -3 V out V in 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Time (s) Electronic neuron. a, Main components of the electronic neuron Fig. 2. proposed here are an analog-to-digital converter (ADC) and a microcontroller. NC means not connected. If the voltage value Vin on the input terminal (I) exceeds a threshold voltage VT (in our experiments VT = 1.5V), the microcontroller connects the input pin to -2.5V and output pin (O) to 2.5V for 10ms, thus sending forward and backward pulses. After that, it waits for a certain amount of time δt and everything repeats again. If Vin < VT , then the microcontroller just continuously samples Vin. The waiting time was selected as δt = τ − γ · (Vin − VT ) + λ (η − 0.5), where τ = 60ms, γ = 50ms/V, λ = 10ms and η is a random number between 0 and 1. In our experimental realization, we used microcontroller dsPIC30F2011 from Microchip with internal 12bits ADC and a possibility of pin multiplexing, so that the only additional elements were two 10k resistors. Actual measurements were done using 5V voltage span, with further assignment of the middle level as zero. The value of resistors shown in a is 10kΩ. b, Response of the electronic neuron on input voltages of different magnitude, the red dashed line is the threshold voltage and is only a guide to the eye. When Vin < VT , no firing occurs. When Vin > VT , the electronic neuron sends pulses, with the average pulse separation decreasing with increasing Vin. Vout is shifted for clarity. lus to a receptor causes receptor potentials whose amplitude is determined by the stimulus strength. When a receptor potential exceeding a threshold value reaches a neuron, the latter starts emitting action potential pulses, whose amplitude is constant but their frequency depends on the stimulus strength. The action potentials are mainly generated along axons in the forward direction. However, there is also a back-propagating part of the signal [28], [29] which is now believed to be responsible for synaptic modifications (or learning) [29], [30]. We have implemented the above behaviour in our electronic scheme as shown in Fig. 2a using an analog-to-digital con- verter and a microcontroller. The input voltage is constantly monitored and once it exceeds a threshold voltage, both forward and backward pulses are generated whose amplitude is constant (we set it here for convenience at 2.5V), but pulse separation varies according to the amplitude of the input signal. In Fig. 2b we show the response of the electronic neu- ron when three resistors of different values are subsequently connected between the input of the neuron and 2.5V. When the resulting voltage (determined by the external resistor and internal resistor connected to the ground in Fig. 2a) is below the threshold voltage (t < 1s in Fig. 2b), no ”firing” occurs (no change in the output voltage). When the input voltage exceeds the threshold (t > 1s in Fig. 2b), the electronic neuron sends forward- and back-propagating pulses. The pulse separation decreases with increase of the input voltage amplitude as it is evident in Fig. 2b. B. Electronic synapse As electronic synapse we have built a memristor emulator, namely an electronic scheme which simulates the behaviour of any memory-resistor. In fact, our memristor emulator can reproduce the behaviour of any voltage- or current-controlled memristive system. The latter is described by the following relations y (t) = g (x, y, t) u (t) , x = f (x, u, t) , (1) (2) where y(t) and u(t) are input and output variables, such as voltage and current, g(x, u, t) is a generalized response (memresistance R, or memductance G), x is a n-dimensional vector describing the internal state of the device, and f (x, u, t) is a continuous n-dimensional vector function [21], [25]. As Fig. 3a illustrates schematically, our memristor emulator consists of the following units: a digital potentiometer, an analog-to-digital converter and a microcontroller. The A (or B) terminal and the Wiper of the digital potentiometer serve as the external connections of the memristor emulator. The resistance of the digital potentiometer is determined by a code written into it by the microcontroller. The code is calculated by the microcontroller according to Eqs. (1) and (2). The analog- to-digital converter provides the value of voltage applied to the memristor emulator needed for the digital potentiometer code calculation. The applied voltage can be later converted to the current since the microcontroller knows the value of the digital potentiometer resistance. In our experiments, we implemented a threshold model of voltage-controlled memristor previously suggested in our earlier paper [31] and also discussed in Ref. [25]. In this model, the following equations (which are a particular case of Eqs. (1,2)) were used: G = x−1, x = (βVM + 0.5 (α − β) [VM + VT − VM − VT ]) (3) ×θ (x − R1) θ (R2 − x) , (4) where θ(·) is the step function, α and β characterize the rate of memristance change at VM ≤ VT and VM > VT , respectively, VM is a voltage drop on memristor, VT is a V VM b AA W B a 0.2Hz 1Hz 5Hz 0.2 0.1 0.0 ) A m ( I 3 VIN+ ADC VIN- Microcontroller Microcontroller R 0 + V(t) - R 0.5 1.0 -0.1 -0.2 -1.0 -0.5 0.0 V M (V) Fig. 3. Electronic synapse. a, Schematic of the main units of the memristor emulator. The memristor emulator consists of a digital potentiometer, ADC and microcontroller. The digital potentiometer unit represents an element whose resistance is defined by a digital code written in it. Two terminals of this unit (A and W) are the external connection terminals of the memristor emulator. The differential ADC converts the voltage between A and W terminals of the digital potentiometer into a digital value. The microcontroller reads the digital code from ADC and generates (and writes) a code for the digital potentiometer according to predefined functions g(x, u, t) and f (x, u, t) and Eqs. (1-2). These operations are performed continuously. In our circuit, we used a 256 positions 10kΩ digital potentiometer AD5206 from Analog Device and microcontroller dsPIC30F2011 from Microchip with internal 12bits ADC. b, Measurements of memristor emulator response when V (t) = V0 cos(2πωt) with V0 ≃ 2V amplitude is applied to the circuit shown in the inset with R0 = 10kΩ. The following parameters determining the memristor emulator response (see Eqs. (3,4)) were used: α = β = 146kΩ/(V·s), VT = 4V, R1 = 675Ω, R2 = 10kΩ. We noticed that the initial value of RM (in the present case equal to 10kΩ) does not affect the long-time limit of the I-V curves. The signals were recorded using a custom data acquisition system. threshold voltage and R1 and R2 are limiting values of memristance. In Eq. (4), the θ-functions symbolically show that the memristance can change only between R1 and R2. In the actual software implementation, the value of x is monitored at each time step and in the situations when x < R1 or x > R2, it is set equal to R1 or R2, respectively. In this way, we avoid situations when x may overshoot the limiting values by some amount and thus not change any longer because of the step function in Eq. (4). In simple words, the memristance changes between R1 and R2 with different rates α and β below and above the threshold voltage. This activation-type model was inspired by recent experimental results on thin- film memristors [23] and as we discuss below it reproduces synapse plasticity. 18 a 18 b 16 14 12 ) 10 V ( e g a t l o V 8 6 4 2 0 Input 1 "sight of food" Input 2 "sound" V 3, in Output "salivation" Probing Learning Probing Input 1 Input 2 V 3, in Output Learning 16 14 12 ) 10 V ( e g a t l o V 8 6 4 2 0 -2 -2 0 3 6 9 12 Time (s) 15 9.0 9.5 10.0 10.5 11.0 Time (s) 4 11.5 12.0 Fig. 4. Development of associative memory. a, Using electronic neurons and electronic synapses (memristor emulators), we have built an electronic circuit corresponding to the neural network shown in Fig. 1. We have used the following parameters defining the operation of memristor emulators: VT = 4V, α = 0, β = 15kΩ/(V·s). At the initial moment of time, the resistance of S1 was selected equal to R1 = 675Ω (lowest resistance state of memristor) and the resistance of S2 was selected equal to R2 = 10kΩ (highest resistance state of memristor). Correspondingly, in the first probing phase, when Input 1 and Input 2 signals are applied without overlapping, the output signal develops only when Input 1 signal is applied. In the learning phase (see also b for more detailed picture), Input 1 and Input 2 signals are applied simultaneously. According to Hebbian rule, simultaneous firing of input neurons leads to development of a connection, which, in our case, is a transition of the second synapse S2 from high- to low-resistance state. This transition is clearly seen as a growth of certain pulses in the signal V3,in (voltage at the input of third neuron) in the time interval from 10.25s to 11s in b. In the subsequent probing phase, we observe that firing of any input neuron results in the firing of output neuron, and thus an associative memory realization has been achieved. The curves in a and b were displaced for clarity. To test that our emulator does indeed behave as a memristor, we have used the circuit shown in the inset of Fig. 3b, in which an ac voltage is applied to the memristor emulator, R, connected in series with a resistor R0 which was used to determine the current. The obtained current-voltage (I- V) curves, presented in Fig. 3b, demonstrate typical features of memristive systems. For instance, all curves are pinched hysteresis loops passing through (0,0) demonstrating no energy storage property of memristive systems [21], [24], [25]. More- over, the frequency dependence of the curve is also typical for memristive systems: the hysteresis shrinks at low frequencies, when the system has enough time to adjust its state to varying voltage, and at higher frequencies, when the characteristic timescale of system variables change is longer than the period of voltage oscillations. C. Associative memory Using the electronic neurons and electronic synapses de- scribed above, we have built an electronic scheme correspond- ing to the neural network depicted in Fig. 1. In this scheme, we directly connect the memristor terminals to the third neuron input. In such configuration, our network behaves essentially as a linear perceptron [32], [33], [34], although different connection schemes are possible. For example, putting a capacitor between the ground and input of the third neuron, we would obtain an integrate-and-fire model [35]. In such a circuit (and also in perceptron networks with many synapses) it is important to ensure that the current from a synapse does not spread to its neighbors. This can be achieved by placing diodes between the right terminals of synapses in Fig. 1 and the third neuron’s input. For our neural network containing only two synapses the effect of current spreading between synapses is not important. Moreover, we would like to highlight that our neural network is fully asynchronous, in distinction to a scheme suggested by Snider [30] based on a global clock. This makes our approach free of synchronization issues when scaling up and closer to a bio-inspired circuit. An asynchronous memristor-based network was also discussed recently [36]. Fig. 4 demonstrates the associative memory development in the present network. Our experiment consists in application of stimulus signals to the first (”sight of food”) and second (”sound”) neurons, and monitoring of the output signal on the third (”salivation”) neuron. We start from a state when the first synaptic connection is strong (low resistance state of the first memristor) and second synaptic connection is weak (high resistance state of the second memristor). In the first ”probing phase” (t < 9s, Fig. 4) we apply separate non-overlapping stimulus signals to the ”sight of food” and ”sound” neurons. This results in the ”salivation” neuron firing when a stimulus signal is applied to the ”sight of food” neuron, but not firing when a stimulus signal is applied to the ”sound” neuron. Electronically, it occurs because pulses generated by the ”sight of food” neuron exceed the threshold voltage of the ”salivation” neuron (due to a low resistance of the first memristor synapse) while the voltage on the ”salivation” neuron input due to the ”sound” neuron pulses is below the threshold voltage of the ”salivation” neuron. In this phase there is no memristor state change since the first memristor is already in its limiting state (with minimal resistance allowed) and its resistance cannot decrease further, and voltage drop on the second memristor is below its voltage threshold. In the ”learning phase” (9s< t <12s, Fig. 4), stimulus voltages are applied simultaneously to both input neurons, thus generating trains of pulses. The pulses from different neurons are uncorrelated, but sometimes they do overlap, owing to a random component in the pulse separation (see Fig. 2 caption for details). During this phase, in some moments of time, back-propagating pulses from the ”salivation” neuron (due to excitation from the ”sight of food” neuron) overlap with forward propagating pulses from the ”sound” neuron causing a high voltage across the second memristor synapse. As this voltage exceeds the memristor threshold, the second synapse state changes and it switches into a low resistance state. It is important to note that this change is possible when both stimuli are applied together (in other words they correlate). As a result, an association between input stimuli develops and the circuit ”learns” to associate the ”sight of food” signal to the ”sound” signal. Our measurements during the second probing phase (t > 12s, Fig. 4) clearly demonstrate the developed association. It is obvious that, in this phase, any type of stimulus - whether from the ”sight of food” or from the ”sound” neurons - results in the ”salivation” neuron firing. Mention should be made about the memristor evolution function given by Eq. 4. In the present neural network the resistance of the second synapse can only decrease, since only negative or zero voltages can be applied to the memristor. A process of synapse depression (corresponding to an increased synapse resistance in our electronic network) can be easily taken into account in different ways. The simplest way is to add a small positive constant to the first line of Eq. 4 (in the parentheses) for the conditioned memristor (S2 in Fig. 1). Then, even at zero value of VM , the resistance of the memristor will slowly increase. Another way is to apply a small positive bias to the conditioned memristor (S2) in the circuit without any changes in Eq. 4 (this would also require taking a non-zero α). III. CONCLUSION We have shown that the electronic (memristive) synapses and neurons we have built can represent important function- alities of their biological counterparts, and when combined 5 together in networks - specifically the one represented in Fig. 1 of this work - they give rise to an important function of the brain, namely associative memory. It is worth again mentioning that, although other memristors (e.g., those built from oxide thin films [37]) could replace the emulator we have built, the electronic neurons and synapses proposed here are electronic schemes that can be built from off-the- shelf inexpensive components. Together with their extreme flexibility in representing essentially any programmable set of operations, they are ideal to study much more complex neural networks [38] that could adapt to incoming signals and ”take decisions” based on correlations between different memories. This opens up a whole new set of possibilities in reproducing different types of learning and possibly even more complex neural processes. ACKNOWLEDGMENT M.D. acknowledges partial support from the National Sci- ence Foundation (DMR-0802830). REFERENCES [1] J. R. Anderson, Language, memory, and thought. Hillsdale, NJ: Erlbaum, 1976. [2] I. P. Pavlov, Conditioned Reflexes: An Investigation of the Physiological Activity of the Cerebral Cortex (translated by G. V. Anrep). London: Oxford University Press, 1927. [3] J. J. Hopfield, ”Neural networks and physical systems with emergent collective computational properties,” Proc. Nat. Acad. Sci. (USA), vol. 79, pp. 2554-2558, 1982. [4] K. Gurney, An Introduction to Neural Networks. UCL Press (Taylor & Francis group), 1997. [5] L. N. Cooper, ”Memories and memory: a physicist’s approach to the brain,” Int. J. Modern Physics A, vol. 15, pp. 4069-4082, 2000. [6] T. Munakata, Fundamentals of the New Artificial Intelligence: Neural, Evolutionary, Fuzzy and More, 2nd ed. Springer, 2008. [7] N. Lewis and S. Renaud, ”Spiking neural networks ”in silico”: from single neurons to large scale networks”, Systems, Signal and Devices Conference, SSD 2007, Hammamet, Tunisia, March 19-22 2007. [8] B. Widrow, “Rate of adaptation of control systems”, ARS Journal, pp. 1378-1385 (1962). [9] M. Mahowald, and R. Douglas, A silicon neuron, Nature, vol.354, pp. 515-518, 1991. [10] R. Jung, E. J. Brauer, and J.!J. Abbas Real-time interaction between a neuromorphic electronic circuit and the spinal cord, IEEE Trans. on Neural Systems and Rehab. Eng., vol. 9, pp.319326, 2001. [11] G. LeMasson, S. Renaud, D. Debay, and T. Bal, Feedback inhibition controls spike transfer in hybrid thalamic circuits, Nature, vol. 4178, pp. 854-858, 2002. [12] M. Sorensen, S. DeWeerth, G. Cymbalyuk, R. L. Calabrese, Using a hybrid neural system to reveal regulation of neuronal network activity by an intrinsic current, J. Neurosci., vol.24, pp. 5427-5438, 2004. [13] R.J. Vogelstein, U. Malik, G. Cauwenberghs, Silicon spikebased synaptic transceiver, Proceedings of ISCAS04, vol.5, array and address-event pp.385-388, 2004. [14] B. Glackin, T.M. McGinnity, L.P. Maguire, QX Wu, A. Belatreche, A novel approach for the implementation of large scale spiking neural networks on FPGA hardware, Proc. IWANN 2005 Computational Intellin- gence and Bioinspired Systems, pp.552-563, Barcelona, Spain, June 2005. [15] J. Schemmel, K. Meier, E. Mueller, A new VLSI model of neural mi- crocircuits including spike time dependent plasticity, Proc. ESANN2004, pp. 405-410, 2004. [16] G. Indiveri, E. Chicca, and R. Douglas. ”A VLSI array of low- power spiking neurons and bistable synapses with spike-timing dependent plasticity”, IEEE Transactions on Neural Networks, vol. 17, pp. 211-221, 2006. [17] J. V. Arthur and K. Boahen, ”Learning in Silicon: Timing is Every- thing”, Advances in Neural Information Processing Systems, vol. 18, Eds. B. Sholkopf and Y. Weiss, MIT Press, 2006. 6 [18] A. Bofill and A. F. Murray. ”Circuits for VLSI implementation of temporally asymmetric Hebbian learning”, Advances in Neural Informa- tion processing systems, vol. 14, Eds. T. G. Dietterich, S. Becker, and Z. Ghahramani, editors, MIT Press, Cambridge, MA, 2001. [19] B. Linares-Barranco, E. Snchez-Sinencio, A. Rodrguez-Vzquez, and J. L. Huertas, ”A CMOS Analog Adaptive BAM with On-Chip Learning and Weight Refreshing”, IEEE Trans. Neural Networks, vol. 4, pp. 445- 455, 1993. [20] L. O. Chua, ”Memristor - The Missing Circuit Element,” IEEE Trans. Circuit Theory, vol. 18, pp. 507-519, 1971. [21] L. O. Chua and S. M. Kang, ”Memrisive devices and systems,” Proc. IEEE, vol. 64, pp. 209-223, 1976. [22] D. B. Strukov, G. S. Snider, D. R. Stewart, and R. S. Williams, ”The missing memristor found,” Nature (London), vol. 453, pp. 80-83, 2008. [23] J. J. Yang, M. D. Pickett, X. Li, D. A. A. Ohlberg, D. R. Stewart, and R. S. Williams, ”Memristive switching mechanism for metal/oxide/metal nanodevices,” Nature Nanotechnology, vol. 3, pp. 429-433, 2008. [24] M. Di Ventra, Yu. V. Pershin, and L. Chua, ”Putting memory into circuit elements: memristors, memcapacitors and meminductors,” Proc. IEEE, vol. 97, pp. 1371-1372, 2009. [25] M. Di Ventra, Yu. V. Pershin, and L. Chua, ”Circuit elements with memory: memristors, memcapacitors and meminductors,” Proc. IEEE, vol. 97, pp. 1717-1724, 2009. [26] T. Driscoll, H.-T. Kim, B.-G. Chae, M. Di Ventra, and D. N. Basov, ”Phase-transition driven memristive system,” Appl. Phys. Lett. vol. 95, pp. 043503/1-3, 2009. [27] Yu. V. Pershin and M. Di Ventra, ”Spin memristive systems: Spin memory effects in semiconductor spintronics,” Phys. Rev. B, Condens. Matter, vol. 78, pp. 113309/1-4, 2008. [28] G. Stuart, M. Spruston, B. Sakmann, and M. Husser, ”Action potential initiation and backpropagation in neurons of the mammalian CNS,” Trends Neurosci., vol. 20, pp. 125-131, 1997. [29] J. Waters, A. Schaefer, and B. Sakmann, ”Backpropagating action potentials in neurons: measurement, mechanisms and potential functions,” Prog. Biophys. Mol. Biol., vol. 87, pp. 145-170, 2005. [30] G. S. Snider, ”Cortical Computing with Memristive Nanodevices,” SciDAC Review, vol. 10, pp. 58-65, 2008. [31] Yu. V. Pershin, S. La Fontaine and M. Di Ventra, ”Memristive model of amoeba’s learning,” Phys. Rev. E, vol. 80, pp. 021926/1-6, 2009. [32] R. Rojas, ”Neural Networks - A Systematic Introduction” (Springer- Verlag, Berlin), 1996. [33] F. Rosenblatt, ”The Perceptron: A Probabilistic Model for Information Storage and Organization in the Brain”, Psych. Rev., vol. 65, pp. 386-408, 1958. [34] M. L. Minsky and S. A. Papert, ”Perceptrons” (Cambridge, MA: MIT Press), 1969. [35] W. Gerstner and W. Kistler, ”Spiking neuron models” (Cambridge university press), 2002. [36] B. Linares-Barranco1 and T. Serrano-Gotarredona, ”Memristance can explain Spike-Time-Dependent-Plasticity in Neural Synapses”, http://hdl.handle.net/10101/npre.2009.3010.1, 2009. [37] J. Borghetti, Z. Li, J. Straznicky, X. Li, D. A. A. Ohlberg, W. Wu, D. R. Stewart, and R. S. Williams, ”A hybrid nanomemristor/transistor logic circuit capable of self-programming,” PNAS, vol. 106, pp. 1699- 1703, 2009. [38] K. Likharev, A. Mayr, I. Muckra, and O. Turel, ”CrossNets: High- Performance Neuromorphic Architectures for CMOL Circuits,” Ann. NY Acad. Sci., vol. 1006, pp. 146-163, 2003.
1808.10315
2
1808
2018-09-04T08:12:34
Deep Learning for Quality Control of Subcortical Brain 3D Shape Models
[ "q-bio.NC" ]
We present several deep learning models for assessing the morphometric fidelity of deep grey matter region models extracted from brain MRI. We test three different convolutional neural net architectures (VGGNet, ResNet and Inception) over 2D maps of geometric features. Further, we present a novel geometry feature augmentation technique based on a parametric spherical mapping. Finally, we present an approach for model decision visualization, allowing human raters to see the areas of subcortical shapes most likely to be deemed of failing quality by the machine. Our training data is comprised of 5200 subjects from the ENIGMA Schizophrenia MRI cohorts, and our test dataset contains 1500 subjects from the ENIGMA Major Depressive Disorder cohorts. Our final models reduce human rater time by 46-70%. ResNet outperforms VGGNet and Inception for all of our predictive tasks.
q-bio.NC
q-bio
Deep Learning for Quality Control of Subcortical Brain 3D Shape Models The ENIGMA Consortium The full author list appears at the end of the paper Abstract. We present several deep learning models for assessing the morphometric fidelity of deep grey matter region models extracted from brain MRI. We test three different convolutional neural net architectures (VGGNet, ResNet and Inception) over 2D maps of geometric features. Further, we present a novel geometry feature augmentation technique based on parametric spherical mapping. Finally, we present an approach for model decision visualization, allowing human raters to see the areas of subcortical shapes most likely to be deemed of failing quality by the ma- chine. Our training data is comprised of 5200 subjects from the ENIGMA Schizophrenia MRI cohorts, and our test dataset contains 1500 subjects from the ENIGMA Major Depressive Disorder cohorts. Our final models reduce human rater time by 46-70%. ResNet outperforms VGGNet and Inception for all of our predictive tasks. Keywords: deep learning, subcortical shape analysis, quality checking 1 Introduction Quality control (QC) has become one of the main practical bottlenecks in big- data neuroimaging. Reducing human rater time via predictive modeling and automated quality control is bound to play an increasingly important role in maintaining and hastening the pace of scientific discovery in this field. Recently, the UK Biobank publicly released over 10,000 brain MRIs (and planning to release 90,000 more); as other biobanking initiatives scale up and follow suit, automated QC becomes crucial. In this paper, we investigate the viability of deep convolutional neural nets for automatically labeling deep brain regional geometry models of failing qual- ity after their extraction from brain MR images. We compare the performance of VGGNet, ResNet and Inception architectures, investigate the robustness of probability thresholds, and visualize decisions made by the trained neural nets. Our data consists of neuroimaging cohorts from the ENIGMA Schizophrenia and Major Depressive Disorder working groups participating in the ENIGMA-Shape project [1]. Using ENIGMAs shape analysis protocol and rater-labeled shapes, we train a discriminative model to separate FAIL(F) and PASS(P) cases. Fea- tures are derived from standard vertex-wise measures. For all seven deep brain structures considered, we are able to reduce human rater time by 46 to 70 percent in out-of-sample validation, while maintaining FAIL recall rates similar to human inter-rater reliability. Our models generalize across datasets and disease samples. Our models' decision visualization, partic- ularly ResNet, appears to capture structural abnormalities of the poor quality data that correspond to human raters' intuition. With this paper, we also release to the community the feature generation code based on FreeSurfer outputs, as well as pre-trained models and code for model decision visualization. 2 Methods Our goal in using deep learning (DL) for automated QC differs somewhat from most predictive modeling problems. Typical two-class discriminative solutions seek to balance misclassification rates of each class. In the case of QC, we focus primarily on correctly identifying FAIL cases, by far the smaller of the two classes (Table 1). In this first effort to automate shape QC, we do not attempt to eliminate human involvement, but simply to reduce it by focusing human rater time on a smaller subsample of the data containing nearly all the failing cases. 2.1 MRI processing and shape features Our deep brain structure shape measures are computed using a previously de- scribed pipeline [2] [3], available via the ENIGMA Shape package. Briefly, struc- tural MR images are parcellated into cortical and subcortical regions using FreeSurfer. Among the 19 cohorts participating in this study, FreeSurfer ver- sions 5.1 and 5.3 were used. The binary region of interest (ROI) images are then surfaced with triangle meshes and spherically registered to a common region- specific template [4]. This leads to a one-to-one surface correspondence across the dataset at roughly 2,500 vertices per ROI. Our ROIs include the left and right thalamus, caudate, putamen, pallidum, hippocampus, amygdala, and nu- cleus accumbens. Each vertex p of mesh model M is endowed with two shape descriptors: Medial Thickness, D(p) = (cid:107)cp− p(cid:107), where cp is the point on the medial curve c closest to p. mapping, J : Tφ(p)Mt → TpM. LogJac(p), Log of the Jacobian determinant J arising from the template Since the ENIGMA surface atlas is in symmetric correspondence, i.e., the left and right shapes are vertex-wise symmetrically registered, we can combine the two hemispheres for each region for the purposes of predictive modeling. Though we assume no hemispheric bias in QC failure, we effectively double our sample. The vertex-wise features above are augmented with their volume-normalized counterparts: {D, J}normed(p) = } . Given discrete area elements of the 3At(p)J(p)D(p). template at vertex p, At(p), we estimate volume as V = (cid:80) , 2 3 {D,J}(p) V { 1 3 We normalize our features subject-wise by this volume estimate to control for subcortical structure size. p∈vrts(M) 2.2 Human quality rating Human-rated quality control of shape models is performed following the ENIGMA- Shape QC protocol (enigma.usc.edu/ongoing/enigma-shape-analysis). Briefly, raters are provided with several snapshots of each region model as well as its placement in several anatomical MR slices. A guide with examples of FAIL (QC=1) and PASS (QC=3) cases is provided to raters, with an additional cat- egory of MODERATE PASS (QC=2) suggested for inexperienced raters. With sufficient experience, the rater typically switches to the binary FAIL/PASS rat- ing. In this work, all QC=2 cases are treated as PASS cases, consistent with ENIGMA shape studies. 2.3 Feature mapping to 2D images Because our data resides on irregular mesh vertices, we first interpolate the fea- tures from an irregular spherical mesh onto an equiangular grid. The interpolated feature maps are then treated as regular 2D images by Mercator projection. Our map is based on the medial curve-based global orientation function (see [5]), which defines the latitude (θ) coordinate, as well as a rotational standardization of the thickness profile D(p) to normalize the longitudinal (φ) coordinate. The resulting map normalizes the 2D appearance of D(p), setting the poles to lie at the ends of the medial curve. In practice, the re-sampling is realized as ma- trix multiplication based on trilinear mesh interpolation, resulting in a 128×128 image for each measure. 2.4 Data augmentation Although our raw sample of roughly 13,500 examples is exceptionally large by the standards of neuroimaging, this dataset may not be large enough to train gen- eralizable CNNs. Standard image augmentation techniques, e.g. cropping and rotations, are inapplicable to our data. To augment our sample of spherically mapped shape features, we sample from a distribution of spherical deforma- tions, i.e. changes in the spherical coordinates of the thickness and Jacobian features. To do this, we first sample from a uniform distribution of vector spher- ical harmonic coefficients Blm, Clm, and apply a heat kernel operator [4] to the generated field on T S2. Change in spherical coordinates is then defined based on the tangential projection of the vector field, as in [4]. The width σ of the heat kernel defines the level of smoothness of the resulting deformation, and the maximum point norm M defines the magnitude. In practice, each random sampling is a composition of a large magnitude, smooth deformation (σ = 10−1, M = 3 × 10−1) and a smaller noisier deformation (σ = 10−2, M = 3 × 10−2). Once the deformation is generated, it is applied to the spherical coordinates of the irregular mesh, and a new sampling matrix is generated, as above. 2.5 Deep learning models We train VGGNet [6], ResNet [7] and Inception [8] architectures on our data. We chose these architectures as they perform well in traditional image classification problems and are well-studied. 2.6 Model decision visualization Deep learning models tend to learn superficial statistical patterns rather than high-level global or abstract concepts [9]. As we plan to provide a tool that both (1) classifies morphometric shapes, and (2) allows a user to visualize what the machine perceives as a 'FAIL', model decision visualization is an important part of our work. Here, we use Prediction Difference Analysis [10], and Grad-CAM [11] to visualize 'bad' and 'good' areas for each particular shape in question. 2.7 Predictive model assessment We use two sets of measures to evaluate the performance of our models. To assess the validity of the models' estimated 'FAIL' probabilities, we calculate the area under the ROC curve (ROC AUC). We also use two supplementary measures: FAIL-recall and FAIL-share. In describing them below, we use the following definitions. TF stands for TRUE FAIL, FF stands for FALSE FAIL, TP stands for TRUE PASS, and FP stands for FALSE PASS. Our first measure, F-recall = T F T F +F P , shows the proportion of FAILS that are correctly labeled by the predictive model with given probability threshold. The second measure, F-share = Number of observations , shows the proportion of the test sample labeled as FAIL by the model. Ideal models produce minimal F-share, and an F-recall of 1 for a given set of parameters. T F +F F 3 Experiments For each of the seven ROIs, we performed three experiments defined by three DL models (VGGNet-, ResNet- and Inception-like architecture). 3.1 Datasets Our experimental data from the ENIGMA working groups is described in Table 1. Our predictive models were trained using 15 cohorts totaling 5218 subjects' subcortical shape models from the ENIGMA-Schizophrenia working group. For a complete overview of ENIGMA-SCZ projects and cohort details, see [12]. To test our final models, we used data from 4 cohorts in the Major Depres- sive disorder working group (ENIGMA-MDD), totaling 1509 subjects, for final out-of-fold testing. A detailed description of the ENIGMA-MDD sites and its research objectives may be found here [13]. Train mean±std 3.4±4.7 max size 16.4 10431 Test mean±std 4.7±4.5 FAIL % accumbens caudate hippocampus thalamus putamen pallidum amygdala 0.8±1.0 0.6±0.6 2.3±3.6 0.9±0.9 3.4 2.6 10436 10436 1.4±1.5 0.4±0.8 1.9±2.0 0.8±0.9 3.5 3018 0.9±0.7 2.0±1.1 2.1 4.2 10433 10436 1.4±1.5 4.9±4.8 3.5 3018 1.5 10436 13.8 10435 10.5 3017 max size 11.4 3018 1.6 3017 3.8 3018 2.1 3018 Table 1: Overview of FAIL percentage mean, standard deviation and maximum for each site. Minimum is equal to 0 for all regions and sites except for hippocam- pus on train (FAIL percentage 5%). Sample sizes for each ROI vary slightly due to FreeSurfer segmentation failure. 3.2 Model validation All experiments were performed separately for each ROI. The training dataset was split into two parts referred to as 'TRAIN GRID' (90% of train data) and 'TRAIN EVAL.' (10% of the data). The two parts contained data from each ENIGMA-SCZ cohort, stratified by the cohort-specific portion of FAIL cases. Each model was trained on 'TRAIN GRID' using the original sampling ma- trix and 30 augmentation matrices resulting in 31x augmented train dataset. We also generated 31 instances of each mesh validation set using each sampling matrix and validated models' ROC AUC on this big validation set during the training. As models produce probability estimates of FAIL (PF AIL), we studied the robustness of the probability thresholds for each model. To do so, we selected PF AIL values corresponding to regularly spaced percentiles values of F-share, from 0.1 to 0.9 in 0.1 increments. For each such value, we examined F-recall the evaluation set. Final thresholds were selected based on the lowest F-share on the TRAIN EVAL set, requiring that F-recall ≥ 0.8, a minimal estimate of inter-rater re- liability. It is important to stress that while we used sample distribution infor- mation in selecting a threshold, the final out-of-sample prediction is made on an individual basis for each mesh. 4 Results Trained models were deliberately set to use a loose threshold for FAIL detection, predicting 0.2-0.5 of observations as FAILs in the 'TRAIN EVAL' sample. These predicted FAIL observations contained 0.85-0.9 of all true FAILs, promising to reduce the human rater QC time by 50-80%. These results largely generalized to the test samples: Table 2 shows our best model and the threshold performance for each ROI. When applied to the test dataset, the models indicated modest over-fitting, with the amount of human effort reduced by 46-70%, while cap- turing 76-94% of poor quality meshes. The inverse relationship between FAIL percentage and F-share (Figure 1) may indicate model failure to learn gener- alizable features on smaller number of FAIL examples. ROC AUC and F-recall performance generalize across the test sites. Since 68% of our test dataset is comprised of the Munster cohort, it is important that overall test performance is not skewed by it. ROI Model Eval AUC Test AUC Eval Test Eval Test F-share F-share F-recall F-recall ResNet 0.8 ResNet 0.9 0.3 Accumbens ResNet 0.86 0.8 0.75 0.5 Amygdala 0.84 0.2 Caudate Hippocampus ResNet 0.85 0.93 0.3 ResNet 0.86 0.91 0.3 Pallidum ResNet 0.88 0.7 Putamen 0.3 0.87 0.4 Thalamus ResNet 0.8 0.35 0.54 0.3 0.36 0.32 0.52 0.47 0.83 0.92 0.82 0.81 0.81 0.93 0.82 0.78 0.8 0.78 0.92 0.91 0.76 0.94 Table 2: Test performance of the best models for each region. ResNet performs the best in all cases. Overall models' performance generalizes to out-of-sample test data. Our experiments with decision visualization (see Fig. 2) indicate that in most FAIL cases, the attention heat map generated by Grad-CAM corresponds to human raters' intuition while Prediction Difference Analysis tend to concentrate on local 'bumps' on shapes. 5 Discussion and Conclusion We have presented potential deep learning solutions for semi-automated quality control of deep brain structure shape data. We believe this is the first DL ap- proach for detecting end-of-the-pipeline feature failures in deep brain structure geometry. We showed that DL can robustly reduce human visual QC time by 46-70% for large-scale analyses, for all seven regions in question, across diverse MRI datasets and populations. Qualitative analysis of models decisions shows promise as a potential training and heuristic validation tool for human raters. There are several limitations of our work. Our planar projection of vertex- wise features introduces space-varying distortions and boundary effects that can affect training, performance and visualization. Recently proposed spherical con- volutional neural nets [14] may be useful to fix this issue. Second, our models' Fig. 1: Scatter plots of F-recall vs. ROC AUC on test datasets and F-share vs. proportion of predicted FAIL cases on test datasets (F-share). Left: F-recall vs ROC AUC. Right: Fail F-share vs FAIL percentage. F-share was calculated based on thresholds from Table 2. Mark size shows the dataset size. Mark shape represents dataset (site): (cid:13) - CODE-Berlin (N=176); (cid:3) - Munster(N=1033); (cid:52) - Stanford (N=105); (cid:53) - Houston(N=195) . Fig. 2: QC report for human raters (left) and decision visualization example based on Grad-CAM for the ResNet model (right). Red colors correspond to points maximizing the model's FAIL decision in the last layer. Decision visu- alization corresponds to the observable deviations from underlying anatomical boundaries indicative of a "FAIL" rating according to an experienced rater. . decision visualization only partly matches with human raters' intuition. In some cases, our models do not consider primary "failure" areas, as assessed by a human rater. Finally, our models are trained on purely geometrical features and do not include information on shape boundaries inside the brain. In rare cases, human 0.60.70.80.91.0F-recall60708090100ROC AUCAccumbensAmygdalaCaudateHippocampusPallidumPutamenThalamus0.00.20.40.60.8F-share024681012FAIL %AccumbensAmygdalaCaudateHippocampusPallidumPutamenThalamus raters pass shapes with atypical geometry because their boundaries look reason- able, and conversely mark normal-appearing geometry as failing due to poor a fit with the MR image. Incorporating intensity as well as geometry features will be the focus of our future work. References 1. Gutman, B., Ching, C., Andreassen, O., Schmaal, L., Veltman, D., Van Erp, T., Turner, J., Thompson, P.M., et al.: Harmonized large-scale anatomical shape anal- ysis: Mapping subcortical differences across the ENIGMA Bipolar, Schizophrenia, and Major Depression working groups. Biological Psychiatry 81(10) (2017) S308 2. Gutman, B.A., Jahanshad, N., Ching, C.R., Wang, Y., Kochunov, P.V., Nichols, T.E., Thompson, P.M.: Medial demons registration localizes the degree of ge- netic influence over subcortical shape variability: An n= 1480 meta-analysis. In: Biomedical Imaging (ISBI), 2015 IEEE 12th International Symposium on, IEEE 1402 -- 1406 3. Roshchupkin*, G.V., Gutman*, B.A., et al.: Heritability of the shape of subcortical brain structures in the general population. Nature Communications 7 (2016) 13738 4. Gutman, B.A., Madsen, S.K., Toga, A.W., Thompson, P.M.: 24. In: A Family of Fast Spherical Registration Algorithms for Cortical Shapes. Volume 8159 of Lecture Notes in Computer Science. Springer International Publishing (2013) 246 -- 257 5. Gutman, B.A., Yalin, W., Rajagopalan, P., Toga, A.W., Thompson, P.M.: Shape matching with medial curves and 1-d group-wise registration. In: Biomedical Imag- ing (ISBI), 2012 9th IEEE International Symposium on. 716 -- 719 6. Simonyan, K., Zisserman, A.: Very deep convolutional networks for large-scale image recognition. arXiv preprint arXiv:1409.1556 (2014) 7. He, K., Zhang, X., Ren, S., Sun, J.: Deep residual learning for image recognition. In: Proceedings of the IEEE conference on computer vision and pattern recognition. (2016) 770 -- 778 8. Szegedy, C., Ioffe, S., Vanhoucke, V., Alemi, A.A.: Inception-v4, inception-resnet and the impact of residual connections on learning. In: AAAI. Volume 4. (2017) 12 9. Jo, J., Bengio, Y.: Measuring the tendency of cnns to learn surface statistical regularities. arXiv preprint arXiv:1711.11561 (2017) 10. Zintgraf, L.M., Cohen, T.S., Adel, T., Welling, M.: Visualizing deep neural network decisions: Prediction difference analysis. arXiv preprint arXiv:1702.04595 (2017) 11. Selvaraju, R.R., Cogswell, M., Das, A., Vedantam, R., Parikh, D., Batra, D.: Grad- cam: Visual explanations from deep networks via gradient-based localization. See https://arxiv. org/abs/1610.02391 v3 7(8) (2016) 12. van Erp, T.G.M., Hibar, D.P., et al.: Subcortical brain volume abnormalities in 2028 individuals with schizophrenia and 2540 healthy controls via the ENIGMA consortium. Molecular psychiatry (2015) 13. Schmaal, L., Hibar, D.P., et al.: Cortical abnormalities in adults and adolescents with major depression based on brain scans from 20 cohorts worldwide in the enigma major depressive disorder working group. Mol Psychiatry 22(6) (2017) 900 -- 909 14. Cohen, T.S., Geiger, M., Koehler, J., Welling, M.: Spherical cnns. arXiv preprint arXiv:1801.10130 (2018) Full author list Dmitry Petrov1,2, Boris A. Gutman2,3, Egor Kuznetsov45, Theo G.M. van Erp4, Jessica A. Turner6, Lianne Schmaal24,25, Dick Veltman25, Lei Wang5, Kathryn Alpert5, Dmitry Isaev1, Artemis Zavaliangos-Petropulu1, Christopher R.K. Ching1, Vince Calhoun40, David Glahn7, Theodore D. Satterthwaite8, Ole Andreas Andreassen9, Stefan Borgwardt10, Fleur Howells11, Nynke Groenewold11, Aristotle Voineskos11, Joaquim Radua13,34,35,36, Steven G. Potkin4, Benedicto Crespo-Facorro14,38, Diana Tordesillas-Gutirrez14,38, Li Shen15, Irina Lebedeva16, Gianfranco Spalletta17, Gary Donohoe18, Peter Kochunov19, Pedro G.P. Rosa20,33, Anthony James21, Udo Dannlowski26, Bernhard T. Baune31, Andr Aleman32, Ian H. Gotlib27, Henrik Walter28, Martin Walter29,41,42, Jair C. Soares30, Stefan Ehrlich43, Ruben C. Gur8, N. Trung Doan9, Ingrid Agartz9, Lars T. Westlye9,37, Fabienne Harrisberger10, Anita Riecher-Rossler10, Anne Uhlmann11, Dan J. Stein11, Erin W. Dickie12, Edith Pomarol-Clotet13,34, Paola Fuentes-Claramonte13,34, Erick Jorge Canales-Rodrguez13,34,39,46, Raymond Salvador13,34, Alexander J. Huang4, Roberto Roiz-Santiaez14,38, Shan Cong15, Alexander Tomyshev16, Fabrizio Piras17, Daniela Vecchio17, Nerisa Banaj17, Valentina Ciullo17, Elliot Hong19, Geraldo Busatto20,33, Marcus V. Zanetti20,33, Mauricio H. Serpa20,33, Simon Cervenka22, Sinead Kelly23, Dominik Grotegerd26, Matthew D. Sacchet27, Ilya M. Veer28, Meng Li29, Mon-Ju Wu30, Benson Irungu30, Esther Walton43,44, and Paul M. Thompson1, for the ENIGMA consortium 1 Imaging Genetics Center, Stevens Institute for Neuroimaging and Informatics, University of Southern California, Marina Del Rey, CA, USA 2 The Institute for Information Transmission Problems, Moscow, Russia 3 Department of Biomedical Engineering, Illinois Institute of Technology, Chicago, IL, USA 4 Department of Psychiatry and Human Behavior, University of California Irvine, Irvine, CA, USA 5 Department of Psychiatry, Northwestern University, Chicago, IL, USA 6 Psychology Department & Neuroscience Institute, Georgia State University, 7 Yale University School of Medicine, New Haven, CT, USA 8 Department of Psychiatry, University of Pennsylvania School of Medicine, Atlanta GA, USA Philadelphia, PA, USA 9 CoE NORMENT, KG Jebsen Centre for Psychosis Research, Division of Mental Health and Addiction, Oslo University Hospital & Institute of Clinical Medicine, University of Oslo, Oslo, Norway 10 Department of Psychiatry, University of Basel, Basel, Switzerland 11 MRC Unit on Risk & Resilience to Mental Disorders, Department of Psychiatry and Mental Health, University of Cape Town, Cape Town, South Africa 12 Centre for Addiction and Mental Health, Toronto, Canada 13 FIDMAG Germanes Hospitalaries Research Foundation, Barcelona, Spain 14 University Hospital Marqus de Valdecilla, IDIVAL, Department of Psychiatry, School of Medicine, University of Cantabria, Santander, Spain 15 Department of Radiology and Imaging Sciences, Indiana University School of Medicine, Indianapolis, IN, USA 16 Mental Health Research Center, Moscow, Russia 17 Laboratory of Neuropsychiatry, Santa Lucia Foundation IRCCS, Rome, Italy 18 School of Psychology, NUI Galway, Galway, Ireland 19 Maryland Psychiatric Research Center, University of Maryland School of 20 Department of Psychiatry, Faculty of Medicine, University of Sao Paulo, Sao Medicine, Baltimore 21 University of Oxford, Oxford, United Kingdom Paulo, Brazil 22 Centre for Psychiatry Research, Department of Clinical Neuroscience, Karolinska 23 Beth Israel Deaconess Medical Center, Harvard Medical School, Boston, MA, USA 24 Orygen, The National Centre of Excellence in Youth Mental Health, Melbourne, Institutet, Stockholm, Sweden 25 Department of Psychiatry, VU University Medical Center, Amsterdam, The Australia Netherlands 26 Department of Psychiatry and Psychotherapy, University of Munster, Germany 27 Department of Psychology, Stanford University, Stanford, CA, USA 28 Charit´e Universitatsmedizin Berlin, corporate member of Freie Universitat Berlin, Humboldt-Universitat zu Berlin, and Berlin Institute of Health, Department of Psychiatry and Psychotherapy CCM, Berlin, Germany 29 Clinical Affective Neuroimaging Laboratory, Leibniz Institute for Neurobiology, Magdeburg, Germany 30 University of Texas Health Science Center at Houston, Houston, TX, USA 31 Discipline of Psychiatry, Adelaide Medical School, The University of Adelaide 32 Interdisciplinary Center Psychopathology and Emotion regulation (ICPE), Neuroimaging Center (BCN-NIC), University Medical Center Groningen, University 33 Center for Interdisciplinary Research on Applied Neurosciences (NAPNA), of Groningen, Groningen, The Netherlands 34 CIBERSAM, Centro Investigacin Biomdica en Red de Salud Mental, Barcelona, University of So Paulo, So Paulo, Brazil Spain 35 Department of Clinical Neuroscience, Centre for Psychiatric Research, Karolinska Institutet, Stockholm, Sweden 36 Department of Psychosis Studies, Institute of Psychiatry, Psychology and Neuroscience, King's College London, United Kingdom 37 Department of Psychology, University of Oslo, Oslo, Norway 38 CIBERSAM, Centro Investigacin Biomdica en Red Salud Mental, Santander, Spain 39 Radiology department, University Hospital Center (CHUV), Lausanne, 40 The Mind Research Network, Albuquerque, NM, USA 41 Leibniz Institute for Neurobiology, Magdeburg, Germany Switzerland 42 Department of Psychiatry and Psychotherapy, University of Tbingen, Tbingen, Germany 43 Division of Psychological and Social Medicine and Developmental Neurosciences, Faculty of Medicine, TU Dresden, Germany 44 Psychology Department, Georgia State University, Atlanta, GA, USA 45 Skolkovo Institute of Science and Technology, Moscow, Russia 46 Signal Processing Laboratory 5 (LTS5), ´Ecole Polytechnique F´ed´erale de Lausanne (EPFL), Lausanne, Switzerland
1011.2797
3
1011
2012-02-01T16:14:01
When are microcircuits well-modeled by maximum entropy methods?
[ "q-bio.NC", "cond-mat.dis-nn", "cs.IT", "cs.IT", "physics.data-an" ]
Describing the collective activity of neural populations is a daunting task: the number of possible patterns grows exponentially with the number of cells, resulting in practically unlimited complexity. Recent empirical studies, however, suggest a vast simplification in how multi-neuron spiking occurs: the activity patterns of some circuits are nearly completely captured by pairwise interactions among neurons. Why are such pairwise models so successful in some instances, but insufficient in others? Here, we study the emergence of higher-order interactions in simple circuits with different architectures and inputs. We quantify the impact of higher-order interactions by comparing the responses of mechanistic circuit models vs. "null" descriptions in which all higher-than-pairwise correlations have been accounted for by lower order statistics, known as pairwise maximum entropy models. We find that bimodal input signals produce larger deviations from pairwise predictions than unimodal inputs for circuits with local and global connectivity. Moreover, recurrent coupling can accentuate these deviations, if coupling strengths are neither too weak nor too strong. A circuit model based on intracellular recordings from ON parasol retinal ganglion cells shows that a broad range of light signals induce unimodal inputs to spike generators, and that coupling strengths produce weak effects on higher-order interactions. This provides a novel explanation for the success of pairwise models in this system. Overall, our findings identify circuit-level mechanisms that produce and fail to produce higher-order spiking statistics in neural ensembles.
q-bio.NC
q-bio
When Do Microcircuits Produce Beyond-Pairwise Correlations? Andrea K. Barreiro1,4,∗, Julijana Gjorgjieva3,5, Fred Rieke2, and Eric Shea-Brown1 1 Department of Applied Mathematics, University of Washington 2Department of Physiology and Biophysics, University of Washington 3 Department of Applied Mathematics and Theoretical Physics, University of Cambridge 4 present affiliation: Department of Mathematics, Southern Methodist University 5 present affiliation: Center for Brain Science, Harvard University ∗ corresponding author Abstract Describing the collective activity of neural populations is a daunting task: the number of possible patterns grows exponentially with the number of cells, resulting in practically unlimited complexity. Recent empirical studies, however, suggest a vast simplification in how multi-neuron spiking occurs: the activity patterns of some cir- cuits are nearly completely captured by pairwise interactions among neurons. Why are such pairwise models so successful in some instances, but insufficient in others? Here, we study the emergence of higher-order interactions in simple circuits with different architectures and inputs. We quantify the impact of higher-order interactions by com- paring the responses of mechanistic circuit models vs. "null" descriptions in which all higher-than-pairwise correlations have been accounted for by lower order statistics, known as pairwise maximum entropy models. We find that bimodal input signals produce larger deviations from pairwise predic- tions than unimodal inputs for circuits with local and global connectivity. Moreover, recurrent coupling can accentuate these deviations, if coupling strengths are neither too weak nor too strong. A circuit model based on intracellular recordings from ON parasol retinal ganglion cells shows that a broad range of light signals induce unimodal inputs to spike generators, and that coupling strengths produce weak effects on higher- order interactions. This provides a novel explanation for the success of pairwise models in this system. Overall, our findings identify circuit-level mechanisms that produce and fail to produce higher-order spiking statistics in neural ensembles. Author Summary Neural populations can, in principle, produce an enormous number of distinct multi-cell patterns -- a number so large that the frequency of these patterns could never be mea- 1 sured experimentally. Remarkably, the activity of many circuits is well captured by simpler probability models that rely only on the activity of single neurons and neuron pairs. These pairwise models remove higher-order interactions among groups of more than two cells in a principled way. Pairwise models succeed even in cases where circuit architecture and input signals seem likely to create a more complex set of outputs. We develop a general approach to understanding which network architectures and input signals will lead such models to succeed, and which will lead them to fail. As a specific application, we consider the remarkable empirical success of pairwise mod- els in capturing the activity of a class of retinal ganglion cells -- the output cells of the retina. Our theory provides a direct explanation for these findings based on the filtering and spike generation properties of ON parasol retinal circuitry, in which collective activity arises through common feedforward inputs to spiking cells together with relatively weak gap junction coupling. Specifically, filtering of light input upstream of ON parasol cells shapes inputs to these cells in such a way that when they are processed by parasol cells, the output spiking patterns are closely fit by a pairwise model. Introduction Information in neural circuits is often encoded in the activity of large, highly interconnected neural populations. The combinatoric explosion of possible responses of such circuits poses major conceptual, experimental, and computational challenges. How much of this potential complexity is realized? What do statistical regularities in population responses tell us about circuit architecture? Can simple circuit models with limited interactions among cells capture the relevant information content? These questions are central to our understanding of neural coding and decoding. Two developments have advanced studies of synchronous activity in recent years. First, new experimental techniques provide access to responses from the large groups of neurons necessary to adequately sample synchronous activity patterns [1]. Second, maximum en- tropy approaches from statistical physics have provided a powerful approach to distinguish genuinely higher-order synchrony (interactions) from that explainable by pairwise statistical interactions among neurons [2 -- 4]. These approaches have produced diverse findings. In some instances, activity of neural populations is extremely well described by pairwise interactions alone, so that pairwise maximum entropy models provide a nearly complete description [5,6]. In other cases, while pairwise models bring major improvements over independent descrip- tions, it is not clear that they fully capture the data [3,7 -- 12]. Empirical studies indicate that pairwise models can fail to explain the responses of spatially localized triplets of cells [11,13], as well as the activity of populations of ∼100 cells responding to natural stimuli [13]. Over- all, the range of empirical results highlights the need to understand the network and input features that control the statistical complexity of synchronous activity patterns. Several themes have emerged from efforts to link the correlation structure of spiking activ- ity to circuit mechanisms using generalized [14 -- 17] and biologically-based models [3, 18, 19]. Two findings are particularly relevant for the present study. First, thresholding nonlinearities 2 in circuits with Gaussian input signals can generate correlations that cannot be explained by pairwise statistics [14]; the deviations from pairwise predictions are modest at moder- ate population sizes [16], but may become severe as population size N grows large [14, 20]. Cluster sizes (i.e., the number of cells firing simultaneously), in particular, may be poorly fit by pairwise models. For large population sizes, a widespread distribution of cluster sizes requires interactions of all orders [14], although for population sizes explored in experimental data third or fourth order interactions may be sufficient [10]. Poor fitting of cluster sizes by the pairwise model is also noted in networks of recurrent integrate-and-fire units with adapt- ing thresholds and refractory potassium currents [19]. Small groups of cells that perform logical operations can be shown to generate higher-order interactions by introducing noisy processes with synergistic effects [4], but it is unclear what neural mechanisms might produce similar distributions. These diverse findings point to the important role that circuit features and mechanisms -- input statistics, input/output relationships, and circuit connectivity -- may play in regulating higher-order interactions. Nevertheless, we still lack a systematic understanding that links these features and their combinations to the success and failure of pairwise statistical models. Second, perturbation approaches can explain why maximum entropy models with purely pairwise interactions capture circuit behavior when the population firing rate is low (i.e. the total number of firing events from all cells in the same small time window is small) [17,21,22]. The fraction of multi-information [4] captured by the pairwise model, a common metric of success, is necessarily low in this regime [17]. In this regime, as well, higher-order interactions cannot be introduced as an artifact of under-sampling the network [22]. The studies which find that pairwise models are successful in capturing multivariate spiking data, however, include cases which either extend beyond [5, 6], or are marginally within [7], the low spikes per time bin regime. Nor do perturbation results necessarily account for the success of models where pairwise connections are restricted to nearest-neighbor connections [5]. Therefore, an explanation for the empirical successes of pairwise models remains incomplete. Here, we aim to bridge some of the gaps between our understanding of mechanistic and statistical models. Our strategy is to systematically characterize the ability of pairwise max- imum entropy (PME) models to capture the responses of several simple circuits (see Figure 1). In each case, we search exhaustively over the entire parameter space for networks with a simple thresholding model of spike generation. Studies of feedforward circuits reveal that the success of PME models does not always bear a simple relation to network architecture -- we find examples in which networks with local connections can deviate substantially from PME predictions, while networks with global connections can be well approximated by PME models. A more consistent determinant of the success of PME models is the unimodal vs. bimodal profile of inputs to the circuit. In addition to departures driven by these inputs, we show that excitatory recurrent coupling can increase departures from PME models by a further 3-5 fold. We apply these general results to responses of specific networks based on measured prop- erties of primate ON parasol ganglion cells. We find that these responses are closely ap- proximated by PME models for a wide range of input types. PME models are successful in 3 this case because the temporal filtering properties of parasol cells induce unimodal synaptic inputs for a broad range of light inputs and because the measured coupling strengths are insufficient to produce large effects on higher-order interactions. This provides insight into why the measured activity patterns in these cells are well captured by PME models [5, 6]. Results Our aim is to determine how network architecture and input statistics influence the success of PME models. We start by considering networks of three cells, which both allow us to make substantial analytical progress and to visualize the results geometrically. We then extend these results in two ways: (1) to models based on the measured properties of primate ganglion cells, one system in which PME models have been very successful [5, 6]; and (2) to larger networks. A geometric approach to identifying higher-order interactions among triplets of cells One strategy to identify higher-order interactions is to compare multi-neuron spike data against a description in which any higher-order interactions have been removed in a principled way -- that is, a description in which all higher-order correlations are completely described by lower-order statistics. Such a description may be given by a maximum entropy model [2, 23, 24], which determines how much of the potential complexity of response patterns produced by large neural populations can be captured by a given set of constraints. The idea is to identify the most unstructured, or maximum entropy, distribution consistent with the constraints. Comparing the predicted and measured probabilities of different responses tests whether the constraints used are sufficient to explain the network activity, or whether additional constraints need to be considered. Such additional constraints would produce additional structure in the predicted response distribution, and hence lower the entropy. A common approach is to limit the constraints to a given statistical order -- for example, to consider only the first and second moments of the distributions, which are determined by the mean and pairwise interactions. In the context of spiking neurons, we denote µi ≡ E[xi] as the firing rate of neuron i and ρij ≡ E[xixj] as the joint probability that neurons i and j will fire. The distribution with the largest entropy for a given µi and ρij is often referred to as the pairwise maximum entropy (PME) model. The problem is made simpler if we consider only permutation-symmetric spiking patterns, in which the firing rate and correlation do not depend on the identity of the cells; i.e. µi = µ, ρij = ρ for i (cid:54)= j. Thus the PME problem is to identify the distribution that maximizes the response entropy given the constraints µ and ρ. In this section, we review some results that help provide a geometric, and hence visual, approach to this problem (see, for example [19]). We consider a permutation-symmetric network of three cells with binary responses. We assume that the response is stationary and uncorrelated in time. From symmetry, the pos- 4 sible network responses are p0 = P (0, 0, 0) p1 = P (1, 0, 0) = P (0, 1, 0) = P (0, 0, 1) p2 = P (1, 1, 0) = P (1, 0, 1) = P (0, 1, 1) p3 = P (1, 1, 1), where pi denotes the probability that a particular set of i cells spike and the remaining 3− i do not. Possible values of (p0, p1, p2, p3) are constrained by the fact that P is a probability distribution, meaning that the sum of pi over all eight states is one. We will rearrange these response probabilities to define a more convenient coordinate system below. Possible solutions to the PME problem take the form of exponential functions character- ized by two parameters, λ1 and λ2, which serve as Lagrange multipliers for the constraints, P (x1, x2, x3) = 1 Z exp[λ1(x1 + x2 + x3) + λ2(x1x2 + x2x3 + x1x3)]. (1) The factor Z normalizes P to be a probability distribution. We can combine the individual probabilities of events p0 = p1 = p2 = p3 = 1 Z 1 Z 1 Z 1 Z exp(λ1) exp(2λ1 + λ2) exp(3λ1 + 3λ2) (cid:18) p2 (cid:19)3 p1 to yield the equation p3 p0 = . (2) This is equivalent to the condition that the strain measure defined in [25] be zero (in particu- lar, the strain is negative whenever p3/p0 − (p2/p1)3 < 0, a condition identified in [25] as cor- responding to sparsity in the neural code). Equation 2 defines, implicitly, a two-dimensional surface in the three-dimensional space of possible probability distributions which we call the maximum entropy surface. The family of distributions consistent with a given µ and ρ forms a line (the iso-moment line) in this space [19]. The PME fit is the intersection of this line with the surface defined by Equation 2. This geometrical description of the PME problem takes a particularly simple form in an alternative coordinate space: fp = p3 + p0 f1p = f1m = p3 p3 + p0 p2 p2 + p1 . 5 (3) This set of coordinates separates network responses based on whether they are "pure" (all cells either spike, or do not) or "mixed" (only a subset of cells spike). fp is the fraction of observed responses that are pure; f1p is the fraction of pure responses with more cells spiking than not (p3 vs. p0). f1m is the fraction of mixed responses with more cells firing than not (p2 vs. p1). Possible probability distributions are contained within a cube in this coordinate space: 0 ≤ fp, f1p, f1m ≤ 1. The iso-moment line for a given µ and ρ is still a line in this coordinate space (see Methods), and the PME approximation is given by the intersection of this line with the maximum entropy surface. The convenience of this coordinate system is apparent when the maximum entropy con- straint (Equation 2) is rewritten: f1p = f 3 1m 1 − 3f1m + 3f 2 1m . (4) This surface is independent of fp -- i.e. the maximum entropy surface forms a curve when projected into the (f1p, f1m)-plane. In addition, each iso-moment line lies in a constant fp plane; the distance of an observed distribution P from the surface is thus easily visualized. This geometric view of the relation between the activity of a given network and its PME fit is particularly useful in visualizing results from exhaustive searches across network parameters, as described below. We use the Kullback-Leibler divergence, DKL(P, P ), to quantify the accuracy of the PME approximation P to a distribution P . This measure has a natural interpretation as the contribution of higher-order interactions to the response entropy S(P ) [2, 4], and may in this context be written as the difference of entropies S( P ) − S(P ). In addition, DKL(P, P ) is approximately − log2 L, where L is the average likelihood -- that is, the average (i.e. per observation) relative likelihood that a sequence of data drawn from the distribution P was instead drawn from the model P [5, 26]. For example, if DKL = 1, the average likelihood that a single sample from P -- i.e. a single network response -- in fact came from P is 2−1 (we use the base 2 logarithm in our definition of the Kullback-Leibler divergence, so all numerical values are in units of bits). An alternative measure of the quality of the pairwise model comes from normalizing DKL(P, P ) by the corresponding distance of the distribution P from an independent max- imum entropy fit DKL(P, P1), where P1 is the highest entropy distribution consistent with the mean firing rates of the cells (equivalently, the independent model P1 is given by the product of single-cell marginal firing probabilities) [2]. Many studies [5 -- 7, 17] use ∆ = Dind − Dpair Dind = 1 − DKL(P, P ) DKL(P, P1) , (5) where following [17] we define Dind ≡ DKL(P, P1) and Dpair ≡ DKL(P, P ). A value of ∆ = 1 (100%) indicates that the pairwise model perfectly captures the additional information left out of the independent model, while a value of ∆ = 0 indicates that the pairwise model gives 6 no improvement over the independent model. Because we are interested in cases where the pairwise model fails to capture circuit outputs, at first it seems we would wish to identify circuits which produce a low ∆. However, in the circuits we explore, we find that the lowest values of ∆ are achieved for nearly independent -- and therefore "uninteresting" -- spike patterns. Therefore, we have chosen to use DKL(P, P ) as our primary measure of the quality of the pairwise model, while reporting values of ∆ in parallel. To get an intuitive picture of DKL(P, P ) throughout the cube of possible distributions P , we view this quantity along constant-fp slices in Figure 2. DKL(P, P ) increases with distance from the constraint curve (Equation 4); along the iso-moment line for a given (µ, ρ), DKL(P, P ) is convex with a minimum of zero at P = P (detailed calculations are given in Materials and Methods). Therefore, for any choice of µ and ρ, DKL(P, P ) increases monotonically as a function of the distance along an iso-moment line. The distance, which is easily visualized, thus gives an indication of how close P comes to being a pairwise maximum entropy distribution. The observed distribution with the maximal deviation from its pairwise maximum entropy approximation will occur at one of the two points where the iso-moment line reaches the boundary of the cube. The global maximum of DKL(P, P ) will, therefore, also occur on the boundary. To assess the numerical significance of DKL(P, P ), we can compare it with the maximal achievable value for any symmetric distribution on three spiking cells. For three cells, this value is 1 (or 1/3 bits per neuron), achieved by the XOR operation [4] -- i.e. the average probability that a single output of the XOR circuit came instead from the PME approxima- tion is 2−1. This distribution, along with its position in the (fp, f1p, f1m) coordinate space, is illustrated in Figure 2 (fp = 0.25 slice) and the right column of Figure 3. We will find that distributions produced by permutation-symmetric networks fall far short of this value. In summary, we have shown that identifying high-order interactions in the joint firing patterns of three cells is equivalent to showing that spiking probabilities lie a substantial distance from a constraint surface that is easy to visualize. Given this geometric description of the problem, we next consider how the distance from the constraint surface depends on circuit connectivity, nonlinear properties of the individual circuit elements, and the statistics of the input signals (Figure 1). When do triplet inputs produce higher-order interactions in spike outputs? We first considered a simple feedforward circuit in which three spiking cells sum and threshold their inputs. Each cell j received an independent input Ij and a "triplet" -- or global -- input Ic that is shared among all three cells. Comparison of the total input Sj = Ic + Ij with a threshold Θ determined whether or not the cell spiked in that time bin. The nonlinear threshold can produce substantial differences between input and output correlations [27 -- 31]. An additional parameter, c, identified the fraction of the total input variance σ2 originating from the global input; that is, c ≡ Var[Ic]/Var[Ic + Ij]. What types of inputs might be natural to consider? The constraint surface (Equations 7 2 and 4) can give us some intuition into what types of inputs will be more or less likely to produce spiking responses that will deviate from the pairwise model. For example, suppose that Ic can take on values that cluster around two separated values, µA < µB, but that it is very unlikely that Ic ever takes on values in the interval between; that is, the distribution of Ic is bimodal. If µB is large enough to push the cells over threshold but µA is not, then we see that any contribution to the right-hand side of Equation 2, p2/p1, depends only on the distribution of the independent inputs Ij; if either one or two cells spike, then the common input must have been drawn from the cluster of values around µA, because otherwise all three cells would have spiked. Keeping the values of µA and µB the same, therefore, but changing the relative likelihood of drawing the common input from one cluster or the other, would change the ratio p3/p0 without changing the ratio p2/p1. Hence the constraint specifying those network responses exactly describable by PME models can be violated when the common input is bimodal. In contrast, we may instead consider a unimodal input -- one whose distribution has a single peak or range of most likely values -- of which a Gaussian input is a natural example. Here, the distribution of the common input Ic is completely described by its mean and variance; both parameters can impact the ratio p3/p0 (by altering the likelihood that the common input alone can trigger spikes) and the ratio p2/p1. Each value of Ic is consistent with both events p1 and p2, with the relative likelihood of each event depending on the specific value of Ic; it is no longer clear how to separate the two. Motivated by these observations, we choose our inputs to come from either one of three types of unimodal distributions -- Gaussian, skewed, or uniform -- or a bimodal distribution. The one-sided skewed shape, in particular, is chosen to mimic qualitative features of inputs to retinal ganglion cells (e.g. [32] and below). We indeed find both qualitative and quantitative differences in the capacity of each class of inputs to generate deviations from the pairwise model. Unimodal inputs fail to produce higher-order interactions in three cell feedfor- ward circuits We first considered unimodal inputs, which were chosen from a distribution with a single peak (or range of most likely values). Gaussian inputs provide a natural example. If Ic and each Ij are Gaussian, then the joint distribution of S = (S1, S2, S3) is multivariate normal, and therefore characterized entirely by its means and covariances. Because the PME fit to a continuous distribution is precisely the multivariate normal that is consistent with the first and second moments, every such input distribution on S exactly coincides with its PME fit. However, even with Gaussian inputs, outputs (which are now in the binary state space {0, 1}3) will deviate from the PME fit [14, 16]. As shown below, non-Gaussian unimodal inputs can produce outputs with larger deviations. Nonetheless, these deviations in all cases are small, and PME models are quite accurate descriptions of circuits with a broad range of unimodal inputs. In more detail, we considered a circuit of three cells with inputs Ic and Ij that could be Gaussian, uniform, or skewed. For each type of input distribution, we probed the output 8 distribution across a range of values for c, σ, and Θ that explored "all" possible activity patterns. In particular, we covered a full range of firing rates (or equivalently spikes per time bin), not limited to the low firing rate regime treated in [17]. Figure 4A-C shows observed distributions for different marginal input statistics (left column). The central col- umn compares all observed distributions with the PME constraint curve, projected into the (f1p, f1m)-plane. The right column of Figure 4A-C shows DKL(P, P ) as a function of c and σ for the value of Θ that maximized DKL(P, P ) (or one of them, if multiple such values exist). Different scales are used to emphasize the structure of the data. For the unimodal cases shown, DKL peaked in regions with comparatively low input variance (σ < 1) and large relative strength of common input (c > 0.5). However, DKL(P, P ) never reached a very high numerical value for unimodal inputs; the maximal values achieved for Gaussian, skewed, and uniform distri- butions are 0.00376 (∆ = 0.989), 0.0152 (∆ = 0.999), and 0.0186 (∆ = 0.9428) respectively (compare with Figure 3). Thus hundreds or thousands of samples would be required to reliably distinguish the outputs of these networks from the PME approximation. Clear patterns emerged when we viewed DKL(P, P ) as a function of output spiking statis- tics rather than input statistics. Figure 5A-B show the same data contained in the center column of Figure 4, but now plotted with respect to the output firing rate, which is the same for all cells. The data were segregated according to the correlation coefficient ρ between the responses of cell pairs, with lighter shades indicating increasing correlation. For a fixed correlation, there was generally a one-to-one relationship between firing rate and DKL(P, P ). For unimodal distributions (Figure 5A-B), DKL(P, P ) showed a double-peaked relationship with firing rate, with larger values attained at low and high firing rates, and a minimum in between. Additionally, DKL(P, P ) had a non-monotonic relationship with spike correlation: it increased from zero for low values of correlation, obtained a maximum for an intermediate value, and then decreased. These limiting behaviors agree with intuition: a spike pattern that is completely uncorrelated can be described by an independent distribution (a special case of PME model), and one that is perfectly correlated can be completely described via (perfect) pairwise interactions alone. Bimodal triplet inputs can generate higher-order interactions in three cell feed- forward circuits Having shown that a wide range of unimodal common inputs produced spike patterns that are well-approximated by PME fits, we next examined bimodal inputs. Figure 4D shows results from a simple ensemble of bimodal inputs -- Bernoulli-distributed common and independent inputs -- that produced moderate deviations from the pairwise approximation. The common input was "1" with probability p and "0" with probability 1 − p. The independent inputs were each chosen to be "1" with probability q and "0" with probability 1− q. The threshold of the cells was between 1 and 2, so that spiking required both common and independent inputs to be active. The space of possible spiking distributions was explored by varying p and q. This circuit produced response distributions that deviated modestly from PME fits, and 9 these distributions preferentially lie on one side of the constraint curve (Figure 4D, center). The largest values of DKL(P, P ) occurred where moderate correlated input is coupled with strong background input (q > 0.5; Figure 4D, right), and reached values that are five times higher than those found for a unimodal distribution (the maximal value achieved is 0.091). The location of the distribution generating this maximum value is demonstrated in the center column of Figure 3. Both of these observations can be explained by direct calculation of the spiking probabili- ties. Substituting the probabilities of different events -- p0 = 1−p+p(1−q)3, p1 = pq(1−q)2, p2 = pq2(1−q) and p3 = pq3 -- into the constraint surface specifying PME models (Equation 2) and dividing by q3, we can write p 1 − p + p(1 − q)3 = 1 (1 − q)3 (6) which gives us an intuition for how to violate the constraint; for a fixed q, we manipulate the left-hand side by changing p. Another way to view this is by observing that the right hand side of Equation 2 can be written without reference to the probability of common input; because P [1 spike Ic = 0] = 0 and P [2 spikes Ic = 0] = 0, one may write p2 p1 = = P [2 spikes Ic = 1]P [Ic = 1] P [1 spike Ic = 1]P [Ic = 1] P [2 spikes Ic = 1] P [1 spike Ic = 1] (7) which has no dependence on the statistics of the common input. So the left hand side of the constraint equation (Eqn. 2) can be manipulated by shifting p, without making any changes to the right hand side. In Figure 5C, we again present values of DKL(P, P ) as a function of the firing rate and pairwise correlation elicited by the full range of possible bimodal inputs. We see that DKL(P, P ) is maximized at a single, intermediate firing rate, and for correlation values near 0.6. We find distinctly different patterns when we view ∆ (i.e. DKL normalized by the distance of P from an independent maximum entropy fit, Equation 5), for these same simulations, as a function of output spiking statistics (Figure 5D-F). For unimodal distributions (Figure 5D- E), ∆ is very close to 1, with the few exceptions at extreme firing rates. For bimodal inputs (Figure 5F), ∆ may be appreciably far from 1 -- as small as 0.5 -- with the smallest numbers (suggesting a poor fit of the pairwise model) occurring for low correlation ρ. This highlights one interesting example where these two metrics for judging the quality of the pairwise model, DKL(P, P ) and ∆, yield contrasting results. An analytical explanation for unimodal vs. bimodal effects We next develop analytical support for the distinct impact of unimodal vs. bimodal inputs on fits by the PME model. Specifically, we calculate DKL(P, P ) for small deviations from the 10 PME constraint surface of Equation 2. We summarize the results of this calculation here; details are in Materials and Methods. We considered narrow distributions of common input Ic, with a small parameter c indicating their variance. By approximating the distribution of network outputs by a Taylor series in c, we found that the leading order behavior of DKL(P, P ) depended on c3 for unimodal distributions -- i.e the low order terms in c dropped out (for symmetric distributions, such as Gaussian, the growth was even smaller: c4). For bimodal distributions, on the other hand, the leading order term of DKL(P, P ) grew like c2. The key point is that, as the strength of common input signals increased, circuits with bimodal inputs diverged from the PME fit much more rapidly than circuits with unimodal inputs. When do pairwise inputs produce higher-order interactions in spike outputs? In the previous sections, we considered permutation-symmetric distributions generated by a single, global, common input. Another class of permutation-symmetric distributions can be generated when common inputs are shared pairwise -- i.e. by two cells but not three at once. We now show that significant departures from the pairwise maximum entropy model (PME) can be generated with pairwise bimodal inputs. This provides a specific example in which network architecture and output statistics are not in simple correspondence. Our circuit setup included three cells, each of which received and summed two inputs and spiked if this sum exceeded a threshold. We denote the inputs I12, I23, I13 so that cell 2 received I12 and I23, and so forth, as illustrated in Figure 1. Each input was chosen from a binary distribution with parameters m and r so that P [Iij = m] = r and P [Iij = 0] = 1 − r. Without loss of generality, we chose m = 1 and chose the threshold such that 1 < Θ < 2. Therefore, both pairwise inputs to a cell must be active in order for a cell to fire. It is not possible in this circuit for precisely two cells to fire; for two cells to fire (say cell 1 and cell 2), both inputs to each cell must be active. However, this implies that both inputs to cell 3 (I13 and I23) are active as well. If two cells fire, then the third must fire as well; that is, p2 = 0. The remaining probabilities are easily computed by itemizing and computing the proba- bilities for each event and are as follows: p3 = r3, p1 = r2(1−r), and p0 = 3r(1−r)2+(1−r)3. This distribution has a unique PME fit consistent with both the first and second moments. However, the PME fit can be far from the actual distribution; we found that DKL(P, P ) de- pended on the input rate r and could exceed 0.5. Thus observation of the network response (single draw from P ) would, on average, have a likelihood ratio of 0.7 of coming from P versus P , and observation of 15 network responses would have a likelihood ratio of less than 0.01. We will revisit the contrast between global and pairwise common inputs below; we also refer to the pairwise case as local inputs. 11 Recurrent coupling produces modest effects on higher-order inter- actions Neural circuits are rarely purely feedforward, and recurrent connectivity can play an impor- tant role in shaping the resulting activity. Therefore, we next modify our thresholding model to incorporate the effects of recurrent coupling among the spiking cells. We take all-to-all coupling among our N = 3 cells, and assume that this coupling is rapid compared with the timescale over which inputs arrive at the cells, so that coupling has its full impact on the spike events that are recorded in a single realization of the model. We limit our study to excitatory interactions, anticipating an application below to the effects of gap junctions known to mediate recurrent coupling between retinal ganglion cells. In more detail, our coupling model may be described as follows: first, inputs arrive at each cell as for the cases without coupling. If the inputs elicit any spikes, there is a second stage in which the input to each neuron receiving a connection from a spiking cell is increased by an amount g. This represents a rapid depolarizing current, assumed for simplicity to add linearly to the input currents. If the second stage results in additional spikes, the process is repeated: recipient cells receive an additional current g, and their summed inputs are again thresholded. The sequence terminates when no new spikes occurred on a given stage; e.g., for N = 3, there are a maximum of three stages. The spike pattern that is recorded on a given trial is the total number of spikes fired across all of the stages. For the simplest case of globally structured inputs and all-to-all coupling, the probability of each spike pattern (or count) can be written down analytically, by evaluating the proba- bilities along the tree of possible events that unfold through the stages just described. For other cases, these trees are more complex, and we instead evaluate probabilities via Monte- Carlo sampling (unimodal inputs), or through enumeration of a finite number of possible input states (bimodal inputs). We begin by describing our results for globally structured, unimodal inputs. We range over parameters c, σ, and Θ as for the uncoupled cases above, and additionally vary g over a wide range of values, from g = 0 to g = 2.4 (comparable to the maximum threshold value). Our basic finding is that coupling has only modest effects on DKL values, when compared with the theoretical range of values that could be obtained. Specifically, the largest value found over all parameters tested was DKL = 0.0099 for the case of Gaussian inputs, 0.108 for uniform, and 0.0599 for skewed. All of these values are roughly 3 − 5× values obtained for the uncoupled cases, and, for Gaussian and skewed inputs, remain much smaller than values found for bimodal inputs. The largest value achieved over any unimodal input and for any coupling strength (0.108) is comparable to the largest value found with bimodal inputs without coupling. Figure 6A shows the data for Gaussian inputs as a scatter plot of DKL values vs. coupling strengths g, showing that intermediate values of coupling -- for which the size of coupling terms is about the same as the standard deviation of the inputs -- have the greatest ability to drive departures from the PME model. Moreover, when coupling strength was varied while keeping other parameters fixed, DKL was also generally maximized at intermediate values of the coupling. In Figure 6B, we show DKL(P, P ) vs. coupling strength g, for representative 12 values of the circuit parameters c, σ and Θ. For each case of global Gaussian, skewed, uniform, and bimodal inputs (thin lines), we see a peak in DKL(P, P ) at medium values of g. We next introduce coupling to circuits with pairwise, locally structured unimodal inputs (as in the preceding section). Here, we range over parameters σ and Θ (c is fixed at 0.5), and additionally vary g over the same wide range of values above. For local inputs, the maximum value of DKL found could be increased modestly by including coupling (from 0.0013 to 0.0086 with Gaussian inputs, from 0.0150 to 0.0157 with skewed inputs, and from 0.0233 to 0.0547 with uniform inputs), but remains substantially below the values found for global inputs with coupling. As in circuits with global inputs, DKL was generally maximized at intermediate values of the coupling; this is illustrated for uniform inputs in the third panel of 6B (note that global and local inputs are identical for Gaussian inputs). An exception was observed for skewed inputs, illustrated in the top panel of 6B, where DKL shows a maximum at g = 0. Finally, we introduce coupling to circuits with bimodal inputs. We note that because the inputs can only take on a finite number of configurations, with all other parameters fixed, the output distribution of the network can only change at a few discrete values of g. Moreover, large values of g induce "all-or-none" firing behavior, which is perfectly matched by pairwise maximum entropy. For global inputs, we find modest increases in DKL which are maximized at an intermediate value of the coupling; for pairwise inputs, the output distribution attains its maximum level of higher order interactions at g = 0 and decreases only when g is sufficiently strong enough to create an "all-or-none" firing pattern (bottom panel, Fig. 6B). In summary, both input statistics and connectivity shape the development of higher- order interactions in our basic circuit model. Other parameters being equal, bimodal inputs can generate a larger DKL(P, P ) than unimodal inputs. For a particular choice of input marginals, global inputs can generate greater deviations than purely pairwise inputs (with the exception of one case, that locally structured bimodal inputs). However, the goodness of the PME fit alone does not distinguish between global and pairwise anatomical projections. Adding fast recurrent excitation increases the accessible magnitude of higher-order interac- tions; for a fixed Θ, the range of accessible DKL increases -- and then decreases -- with the magnitude of recurrent excitation. With these principles in hand, we now study a more biologically realistic network model. An experimentally constrained model for correlated firing in retinal ganglion cells PME approaches have been effective in capturing the activity of small retinal ganglion cell (RGC) populations [5 -- 7]. This success does not have an obvious anatomical correlate -- i.e. there are multiple opportunities in the retinal circuitry for interactions among three or more ganglion cells. Why do these apparently fail to generate higher-order interactions in output spiking data? To answer this question, we explored the properties of circuits composed of cells with input statistics, recurrent connectivity and spike-generating mechanisms based 13 directly on experiment. We based our model on ON parasol RGCs, one of the RGC types for which PME approaches have been applied extensively [5, 6]. We first describe the RGC model, then apply this to feedforward circuits, and finally consider recurrent connections. RGC model We modeled a single ON parasol RGC in two stages (for details see Materials and Methods). First, we characterized the light-dependent excitatory and inhibitory synaptic inputs to cell k (gexc k (t)) in response to randomly fluctuating light inputs s(t) via a linear-nonlinear model, e.g.: k (t), ginh k (t) = N exc[Lexc ∗ sk(t) + ηexc gexc k ], (8) where N exc is a static nonlinearity, Lexc is a linear filter, and ηexc is an effective input noise that captures variability in the response to repetitions of the same time-varying stimulus. These parameters were determined from fits to experimental data collected under condi- tions similar to those in which PME models have been tested empirically. The modeled excitatory and inhibitory conductances captured many of the statistical features of the real conductances, particularly the correlation time and skewness. k Second, we used Equation 8 and an equivalent expression for ginh k (t) as inputs to an integrate-and-fire model incorporating a nonlinear voltage and history-dependent term to account for refractory interactions between spikes [33]. The voltage evolution equation was of the form dV dt = F (V, t − tlast) + Iinput(t) , C (9) where F (V, t − tlast) was allowed to depend on the time of the last spike tlast (more details in Methods). We fit the parameters of this model to a dynamic clamp experiment [34, 35] in which currents corresponding to gexc(t) and ginh(t) were injected into a cell and the resulting voltage response measured. Excitatory and inhibitory synaptic currents injected during one time step were determined by scaling the conductances by driving forces based on the measured voltage in the previous time step. Recurrent connections were implemented by adding an input current proportional to the voltage difference between the two coupled cells. The prescription above provided a flexible model that we used to study the responses of small RGC networks to a wide range of light inputs and circuit connectivities. Specifically, we simulated RGC responses to light stimuli that were (1) constant, (2) time-varying and spatially uniform, and (3) varying in both space and time. Correlations between cell inputs arose from shared stimuli, from shared noise originating in the retinal circuitry [32], or from recurrent connections [32, 36]. Shared stimuli were described by correlations among sk. Shared noise arose via correlations in ηk as described in Materials and Methods. The recurrent connections were chosen to be consistent with observed gap-junctional coupling between ON parasol cells. We also investigated how stimulus filtering by Lexc and Linh influenced network statistics. 14 For each network model, we determined whether the accuracy of a PME fit to the outputs was predictable based on the structure of the input distributions, using the results developed above for the idealized circuit model. We focused on excitatory conductances because they exhibit stronger correlations than inhibitory conductances in ON parasol RGCs [32]. To compare our results with empirical studies, constant light and spatially and temporally fluctuating checkerboard stimuli were used as in [5, 6]. Feedforward RGC circuits k k and ηinh We start by considering networks without recurrent connectivity and without temporal mod- ulations in the light input. Thus we set sk(t) = 0 for k = 1, 2, 3, so that the cells received only Gaussian correlated noise ηexc and constant excitatory and inhibitory conduc- tances. Time-dependent conductances were generated and used as inputs to a simulation of three model RGCs. Simulation length was sufficient to ensure significance of all reported deviations from PME fits (see Materials and Methods). Under these conditions the excita- tory conductances were unimodal and broadly Gaussian. As expected from earlier results on threshold models, the spiking distributions were well-modeled by a PME fit, as shown in Figure 7A; DKL(P, P ) is 2.90 × 10−5 bits. This agrees with the very good fits found experimentally in [5] under constant light stimulation. For full-field simulations, each cell received the same stimulus, sk(t) = s(t), where s(t) refreshes every few milliseconds with an independently chosen value from one of several marginal distributions. The shared stimulus produced strong pairwise correlation between conductances of neighboring cells. However, results from our threshold model (Fig. 4) suggest that this is not the overall determining factor in whether or not spiking outputs will be well-modeled by a PME fit; rather, the shape of the marginal distribution of inputs (here, conductances) should be more important than the strength of pairwise correlations. We first examined the effects of different marginal statistics of light stimuli, standard deviation of full-field flicker, and refresh rate on the marginal distributions of excitatory conductances. For a short refresh rate (8 ms) and small flicker variance (1/6 or 1/4 of baseline light intensity), temporal averaging via the filter Lexc and the approximately linear form of N exc over these light intensities produced a unimodal, modestly skewed distribution of excitatory conductances, regardless of whether the flicker is drawn from a Gaussian or binary distribution (see Figure 7B-C, center panels). For a slower refresh rate (100 ms) and large flicker variance (1/3 or 1/2 of baseline light intensity), excitatory conductances had multi-modal and skewed features, again regardless of whether the flicker is drawn from a Gaussian or binary distribution (Figure 7D). Other parameters being equal, binary light input produced more skewed conductances. While some conductance distributions had mul- tiple local maxima, these were never well-separated, with the envelope of the distribution still resembling a skewed distribution. As expected from our studies with the simple thresholding model of spike generation, the largely unimodal shape of input distributions was reflected in the ability of PME fits to accurately capture spiking distributions. DKL(P, P ) values computed from the observed distributions were small, never exceeding 0.0067; these are numbers comparable to what is 15 achievable by skewed inputs in our simple thresholding circuit model. To test the sensitivity of this conclusion to the finite sampling in our simulations, we performed an analysis in which the data was divided into 20 subsets and the maximum entropy analysis was performed indi- vidually on each subset. The resulting KL-distance remained small, never exceeding 0.0089. In summary, even high-contrast, bimodal, highly spatially correlated stimulus variations do not produce a large departure from the PME fit. When we examined all of the spiking distributions produced in this sequence of simula- tions, we found a common pattern in the way in which the PME fit deviated from observed distributions. Single spiking events were over-predicted by PME fits, whereas double spiking events were under-predicted. We note that this is the same situation observed in our simple threshold model with bimodal global inputs (see Figure 3 and Materials and Methods), and corresponds to the case of negative strain identified by Ohiorhenuan et al. [25]. This find- ing is extremely robust; upon perturbing the distributions by estimated standard errors, as described in Materials and Methods, only 22 out of 840 perturbed distributions showed a positive strain while the remainder had negative strain. Overall, high-pass filtering -- a consequence of the differentiating linear filter in Equation 18 and illustrated in Figure 7D -- was responsible for significantly reducing the bimodality of the input stimuli. In Figure 7E, we produce a filter without the biphasic, high-pass shape of Equation 18 (i.e., without the negative dip at longer time lags) by simply rectifying the experimentally determined filter. This filter produced conductance distributions that more completely reflect the bimodal shape of binary light inputs (Figure 7E, center panel). The resulting simulation produces six-times greater DKL(P, P ) (Figure 7E, right panel). This raises the intriguing suggestion that greater DKL(P, P ) may occur for other cell types primarily characterized via monophasic filters (such as midget cells), or for different light stimuli (e.g., at different mean levels) for which the retinal circuit acts to primarily integrate rather than differentiate over time. In Figure 8A, we examine this effect over all full-field stimulus conditions by plotting deviations from the pairwise model, in terms of DKL(P, P ), in simulations with a rectified filter against the same quantity with a non-rectified filter. An increase in DKL(P, P ) was observed across stimulus conditions, with a markedly higher effect for longer refresh rates. Large effects were accompanied by a striking increase in the bi- or multi-modality of ex- citatory conductances (see Figure 8B, illustrating binary stimulus at 100 ms refresh rate and standard deviation of 1/4), while small effects were accompanied by small or no dis- cernible change in the marginal distribution of excitatory synaptic conductances (see Figure 8C, illustrating binary stimulus at 8 ms refresh rate and standard deviation of 1/2). This consistent change could not be attributed to changes in lower order statistics; there was no consistent relationship between the change in pairwise model performance and either firing rate or pairwise correlations (data not shown). Moving beyond full field light stimuli, we asked whether pairwise maximum entropy models capture RGC responses to stimuli with varying spatial scales. We fixed stimulus dynamics to match the two cases that yielded the highest DKL(P, P ) under the full field protocol: for both Gaussian and binary stimuli, this was an 8 ms refresh rate and σ = 1/2. 16 The stimulus was generated as a random checkerboard with squares of variable size; each square in the checkerboard, or stixel, was drawn independently from the appropriate marginal distribution and updated at the corresponding refresh rate. The conductance input to each RGC was then given by convolving the light stimulus with its receptive field, where the stimulus is positioned with a fixed rotation and translation relative to the receptive fields. This position was drawn randomly at the beginning of each simulation and held constant throughout (see Figure 9D,G for examples, and Materials and Methods for further details). The RGC spike patterns remained very well described by PME models for the full range of spatial scales. Figure 9A shows this by plotting DKL(P, P ) vs. stixel size. Values of DKL(P, P ) increased with spatial scale, sharply rising beyond 128 µm, where a stixel is approximately the same size as a receptive field center. The points at 512 µm are from the corresponding full field simulations, illustrating that introducing spatial scale via stixels produces even closer fits by PME models. Values reported in Figure 9A are averages of DKL(P, P ) produced by 5 random stimulus positions. At stixel sizes of 128 µm and 256 µm, the resulting spiking distributions differed significantly from position to position; in Figure 9B, we show the probabilities of the distinct singlet (e.g. P (1, 0, 0)) and doublet (e.g. P (1, 1, 0)) spiking events produced at 256 µm. Each stimulus position creates a "cloud" of dots (identified by color); large dots show the average over 20 sub-simulations. Each sub-simulation is identified by a small dot of the same color; because the simulations are very well-resolved, most of these are contained within the large dots (and hence not visible here). Heterogeneity across stimulus positioning is indicated by the distinct positioning of differently colored dots. At smaller spatial scales, the process of averaging stimuli over the receptive fields results in spiking distributions that are largely unchanged as stimulus position changes, as shown in Figure 9C, where singlet and doublet spiking probabilities are plotted for 60 µm stixels. Thus, filtered light inputs are largely homogeneous from cell to cell, as each receptive field samples a similar number of independent, statistically identical inputs: Figure 9D shows the projection of input stixels onto cell receptive fields from an example with 60 µm stixels. The resulting excitatory conductances (Figure 9E) and spiking patterns (Figure 9F) are very close to cell-symmetric. By contrast, spiking patterns showed significant heterogeneity from cell to cell,when the stixel size was large; this arises because each cell in the population may be located differently with respect to stixel boundaries, and therefore receive a distinct pattern of input activity. Figure 9G shows the projection of input stixels onto cell receptive fields from one example, with 256 µm stixels. The difference in statistical properties of inputs is reflected in the marginal distributions of excitatory conductances that each cell receives, shown in Figure 9H. It is also apparent in the observed output spiking patterns, where one particular doublet spiking pattern (here, 110) is significantly more likely to occur than the others (see Figure 9I). However, PME models gave excellent fits to data regardless of heterogeneity in RGC responses, as illustrated for this particular example in Figure 9I; over all 20 sub-simulations (as above), and over all individual stixel positions, we found a maximal DKL(P, P ) value of 0.00811. 17 Recurrent connectivity in the RGC circuit We next considered the role of recurrence in shaping higher order interactions by incorpo- rating gap junction coupling into our simulations. We did this separately for each full field stimulus condition: Gaussian and binary marginal distributions, with variances 1/16, 1/12, 1/8, 1/6, 1/4, 1/3, or 1/2 of baseline light intensity, and refreshed at either 8, 40, or 100 ms intervals. In each case, we added gap junction coupling with strengths from 1 to 16 times an experimentally measured value [32], and compared the resulting DKL with that obtained without recurrent coupling. Results are shown in Figure 10. At the measured coupling strength (ggap = 1.1 nS) itself, the fit of the pairwise model barely changed (Figure 10B). At twice the measured coupling strength (ggap = 2.2 nS), recurrent coupling had increased higher order interactions, as measured by larger values of DKL for all tested stimulus conditions. Higher order interactions could be further increased, particularly for long refresh rates (100 ms), by increasing the coupling strength to 4 or 8 times its baseline level (ggap = 8.8 nS; see Figure 10C, D). Consistent with our intuition that very strong coupling leads to "all-or-none" spiking patterns, DKL(P, P ) decreased as ggap increased further, often to a level below what was seen in the absence of coupling (Figure 10F). Again, we considered the possibility that firing rates could explain the increase; however, firing rates actually decreased with the addition of coupling. Our findings are consistent with our results for the threshold model (Figure 6), in that the impact of coupling is maximized at intermediate values of the coupling strength. Moreover, the impact of recurrent coupling on the maximal values of DKL evoked by visual stimuli is small overall, and almost negligible for experimentally measured coupling strengths. Scaling of higher-order interactions with population size The results above identify the input statistics -- particular unimodality vs bimodality -- as a key determinant of the success of PME models. How robust is this conclusion to network size? The permutation-symmetric architectures we have considered can be scaled up to more than three cells in several natural ways; for example, we can consider N cells with a global common input. The pairwise (local) input structure can also be scaled up to consider N cells on a ring, with each pair of adjacent cells receiving a common, pairwise input (see Figure 1). We first considered a sequence of models in which a set of N threshold spiking units received global input Ic (with mean 0 and variance σ2c) and an independent input Ij (with mean 0 and variance σ2(1− c)). As for the three cell networks above, the output of each cell was determined by summing and thresholding these inputs. The probability distribution of network outputs was computed as described in the Methods and then fit with a pairwise maximum entropy distribution. As also for the three cell networks, we explored a range of σ, c, and Θ and recorded the maximum value of DKL(P, P ) between the observed distribution P and its PME fit P . Figure 11 shows this DKL/N (i.e. entropy per cell [16]) for Gaussian, uniform, skewed, and bimodal input distributions. We found that the maximum DKL(P, P ) increased roughly linearly with N for bimodal 18 inputs, and superlinearly for unimodal inputs. The relative ordering found at N = 3 -- that the maximal achievable DKL(P, P ) is lowest for Gaussian inputs, followed by skewed, uniform, and bimodal inputs consecutively -- remained the same. The sidebar of Figure 11 shows that the probability distributions produced by these inputs qualitatively agree with this trend: departures from PME were more visually pronounced for global bimodal inputs (top histogram) than for global unimodal inputs (third histogram from top). At N = 16, the value DKL/N ≈ 0.1 for bimodal global inputs corresponds to a likelihood ratio of 0.33 that a single draw from P (single network output) in fact came from the PME fit P versus P ; a likelihood < 0.01 is reached for 4 draws. We next considered pairwise inputs for N > 3 cells by adopting a ring structure with nearest-neighbor common inputs (illustrated in Figure 1). For unimodal inputs, we computed DKL(P, P ) while varying σ and Θ; for bimodal inputs, we varied the probability of each Bernoulli input. Figure 11 shows the maximal DKL(P, P ) per neuron. Overall, circuits with bimodal pairwise inputs showed appreciable values that are about half of that found for bimodal global inputs. The relatively large deviation at N = 3 receded, replaced by deviations that were similar to those seen for global, unimodal inputs. For pairwise unimodal inputs, values of DKL(P, P ) remained very small. Finally, we extended our recurrent model to N > 3 cells. As for the three cell networks, we explored a range of σ, c, and Θ for unimodal inputs, or s, p, and Θ for bimodal inputs. In addition, the coupling strength, g, was varied for each type of input. Coupling was either all-to-all (global) or applied pairwise in a ring structure (local). Figure 12 shows the maximal DKL(P, P ) per neuron, for each type of input and network architecture, up to population size N = 8. Recurrent coupling increases the available range of higher order interactions in most settings from the level achieved with purely feed forward connections. However, the maximal value of DKL(P, P ) per cell is steady or decreasing with N , suggesting that recurrence has less impact as population size grows. To summarize, the greater impact of bimodal vs. unimodal input statistics on maximal values of DKL(P, P ) that can be obtained in a given circuit persists from N = 3 up to N = 16 (tested up to N = 8 in the recurrent case). Moreover, agreeing with intuition, global inputs of a given type can generate greater deviations than purely pairwise inputs (with the exception of one case, N = 3). Overall (again excepting this one case), for the circuit parameters producing maximal deviations, it becomes easier to statistically distinguish be- tween spiking distributions and their PME fits as N increases in feedforward networks. In contrast, the maximal attainable DKL(P, P ) per cell appears to decrease with N in many recurrent networks. Discussion We use simple mechanistic models to identify which combinations of network architectures and input signals produce spike patterns with higher-order interactions and which do not. Deviations in circuit outputs from pairwise maximum entropy (PME) predictions were much smaller than the maximum theoretically attainable values for a general spiking pattern. 19 Moreover, output statistics were not simply related to network architecture. Nonetheless, several simple principles emerge that determine the strength of higher-order interactions. First, bimodal input distributions produced stronger higher-order interactions than unimodal distributions. Second, networks with shared inputs among all cells produced greater higher- order interactions than those with pairwise inputs (except for the case of three cell networks receiving bimodal input). Third, recurrent excitatory or gap junction coupling could produce a further, moderate increase higher-order correlations; the effect was greatest for coupling of intermediate strength. Our overall results held for networks with nonlinear integrate-and-fire units based on measured properties of retinal ganglion cells. Together with the facts that ON parasol cell filtering suppresses bimodality in light input, and that coupling among ON parasol cells is relatively weak, our findings provide an explanation for why their population activity is well captured by PME models. Comparison with empirical studies How do our maximum entropy fits compare with empirical studies? In terms of DKL(P, P ) -- equivalently, the logarithm of the average relative likelihood that a sequence of data drawn from P was instead drawn from the model P -- numbers obtained from our RGC models are very similar to those obtained by experiments on retinal ganglion cells [5, 6]. We find that DKL(P, P ) = 2.90× 10−5 bits under constant light conditions, compared to an experimental value of 0.0008 [5] (inferred from a reported likelihood ratio of 0.99944). Under full-field, time varying light conditions, as well as spatiotemporally varying stixel simulations, we find average log-likelihood ratios of up to one order of magnitude larger -- bounded above by 0.007. We can view this as a model of the checkerboard experiments of [5], for which close fits by PME distribution were also observed (likelihood numbers were not reported). Similarly, the values of ∆ that are produced by our RGC model are close to those found by [5, 7] under comparable stimulus conditions. We obtain ∆ = 99.5% (for cell group size N = 3) under constant illumination, which is near the range reported by [5] (97.8 − 99.2%, N = 3 − 7). For full-field stimuli we find a range of numbers from 95.7 − 99.3% (N = 3). The simple threshold models that we have developed, meanwhile, give us a roadmap for how circuits could be driven in such a way as to lower ∆. Figure 5D-F show ∆ plotted as a function of firing rate for the data presented in Figure 5A-C: circuits of N = 3 cells receiving global common inputs. We observe that ∆ ≈ 1 for Gaussian and skewed inputs over a broad range of firing rates and pairwise correlation coefficients, but that values of ∆ can be depressed by 10-15% in the presence of a bimodal common input. Indeed, Shlens et al. [5] showed that adding global bimodal inputs to a purely pairwise model can lead to a comparable departure in ∆. Our results are consistent with this finding, and explicitly demonstrate that the bimodality of the inputs -- as well as their global projection -- are characteristics that lead to this departure. While meaningful in an experimental study with non-neglible pairwise correlations, we caution that using ∆ as a metric can be problematic when an idealized circuit is explored over its full range of parameters, because it may flag "uninteresting" cases in which cells are nearly independent, and a pairwise model adds little additional value. Specifically, if Dind 20 is small (i.e., the true distribution is well-approximated by the independent distribution P1), then ∆ may be appreciably far from 1 although Dpair is small. Thus, a poor pairwise maximum entropy fit, as measured by ∆ (that is, ∆ < 1) is not necessarily indicative of a poor performance in DKL(P, P ). For example, in the bimodal common input case (Figure 5F), the very lowest values of ∆ are achieved for low correlation ρ; in essence, when the independent model already does a good job of representing the output distribution. As suspected this performance is not reflected in Figure 5C, where low correlation gives low DKL(P, P ). In summary, ∆ can be as low as 0.5 for distributions that are barely perceptibly different when measured by the Kullback-Leibler divergence. Figure 13A-B extend our observations to a circuit of N = 12 cells forced by Gaussian and skewed inputs respectively. We find that small ∆ occurs for σ (cid:28) 1, coincident with low population firing rates. Here P is nearly independent (because it is dominated by 0 spiking events, the distribution is well-modeled by independent non-spiking neurons) and the improvement of the pairwise model over the independent model is negligible. The low population firing rate regime (N ν < 1, where ν is the single cell firing rate per time bin and N is the population size) is also the regime where 1− ∆ is linear in N − 2 [17]. If a nontrivial deviation of ∆ (from 1) is observed for N ν < 1 (such as, in Figure 5, N = 3), then 1 − ∆ must continue to grow as N grows; equivalently, ∆ must decrease. The growth of 1− ∆ with N for particular points in this region is illustrated in Figure 13C. Using correlation structure to infer anatomical structure We address two questions about the relationship between the architecture of feedforward circuits and the statistical structure of the spike patterns that they produce, based on our comparisons between global-input and pairwise nearest-neighbor network architectures in the Results. First, if a circuit produces spike patterns that deviate substantially from pairwise maximum- entropy (PME) predictions, can we conclude that it has beyond-pairwise anatomical pro- jections -- that is, common inputs received by more than two cells? For a small group of N = 3 cells, the answer is no: we find that, among all cases we study, the largest deviation from PME predictions occurs for purely pairwise (binary) inputs, so that departures from PME models do not imply departures from pairwise nearest-neighbor network architectures. For larger N , the answer is a qualified yes: for marginal statistics of a given type, we show in Figure 11 that the greatest deviations from PME models do correspond to global com- mon (as opposed to purely pairwise) inputs. However, without knowing input marginals, values of DKL are still not predictive of anatomy: for example, if N = 16, then roughly the same values of DKL(P, P ) follow from global inputs with uniform marginals as for pairwise nearest-neighbor inputs with binary marginals. Second, if a circuit produces spike patterns that are well-described by PME models, does this imply that it has a pairwise architecture? Again the answer is no, as the success of PME models depends on N and marginal statistics. For N > 3 and known input marginals, better fits by PME imply pairwise nearest-neighbor connectivity; otherwise, such inferences cannot be made. 21 Scope and open questions Our first set of findings are for a set of circuit models with a simple thresholding nonlinearity at each cell. These models were chosen to be simple enough to allow analytical insights and a complete parametric study. While our retinal ganglion cell model demonstrates that these findings, based on a simple threshold model, do carry over to describe the spiking statistics of a more realistic spiking model (here, a time-dependent, nonlinear integrate-and-fire system), there are many aspects of circuits left unexplored by our study. Prominent among these is heterogeneity. Only a few of our simulations produce heterogeneous inputs to model RGCs, and all of our stud- ies apply to cells with identical response properties. This is in contrast to studies such as [7] which examine correlation structures among multiple cell types. For larger networks, feedforward connections with variable spatial profiles also occur, between the extremes of "nearest neighbor" and global input connections examined here. It is also possible that more complex input statistics could lead to greater higher-order interactions [37]. Finally, Figure 11 indicates that some trends in DKL(P, P ) vs. N appear to become nonlinear for N (cid:39) 10; for larger networks, our qualitative findings could change. Our study also leaves largely open the role of different retinal filters in generating higher- order interactions. We have found that the specific filtering properties of ON parasol cells suppress bimodality in light inputs, suggesting that other classes of retinal ganglion cells may produce more robust higher-order interactions (compare Figures 7D, E). This predicts a specific mechanism for the development of higher-order interactions in preparations that include multiple classes of ganglion cells [7]. Finally, we considered circuits with a single step of inputs and simple excitatory or gap junction coupling; a plethora of other network features could also lead to higher-order interactions, including multilayer feedforward struc- tures, together with lateral and feedback coupling. We speculate that, in particular, such mechanisms could contribute to the higher-order interactions found in cortex [8, 10, 11, 38]. A final outstanding area of research is to link tractable network mechanisms for higher- order interactions with their impact (or lack of impact) on information encoded in neural populations [10, 38]. A simple starting point is to consider rate-based population codes in which each stimulus produces a different "tuned" average spike count (see, e.g., Ch. 3 of [39]). One can then ask whether spike responses can be more easily decoded to estimate stimuli for the full population response (i.e., P ) to each stimulus or for its pairwise approximation ( P ). In our preliminary tests where higher-order correlations were created by inputs with bimodal distributions, we found examples where decoding of P vs. P differed substantially. However, a more complete study would be required before general conclusions about trends and magnitudes of the effect could be made; such a study would include complementary approach in which the full spike responses P are themselves decoded via a "mismatched" decoder based on the pairwise model P [38]. Overall, we hope that the present paper, as one of the first that connects circuit mechanisms to higher-order statistics of spike patterns, will contribute to future research that takes these next steps. 22 Materials and Methods DKL and distance from PME surface We can see that the curve along which µ and ρ are constant -- the iso-moment line -- remains a line in the (fp, f1p, f1m) coordinate space by inverting the equations µ = p1 + 2p2 + p3, ρ = p2 + p3, and 1 = p0 + 3p1 + 3p2 + p3. The first coordinate, fp ≡ p0 + p3 = 3(ρ − µ) − 1, is constant for a fixed (µ, ρ). Similarly, f1m and f1p can be written as linear functions of the remaining free parameter. To see that DKL(P, P ) is convex along an iso-moment line, we consider DKL(P, P ) as P varies so as to remain on an iso-moment line. Letting f1m and f1p be the coordinates of the PME fit, and defining dm = f1m − f1m and dp = f1p − f1p, we find where dx is an increment of distance along the iso-moment line. Inverting Equations 3 and substituting the results into the definition of DKL, we can write (cid:33) (cid:19) 1 − dp 1 − f1p (cid:18) (cid:19) 1 − dm 1 − f1m dp = dm = (cid:113) (cid:113) (cid:32) (cid:18) dp f1p (cid:18) dx dx (fp − 1)/3 p + (1 − fp)2/9 f 2 fp p + (1 − fp)2/9 f 2 (cid:32) (cid:33) (cid:19) 1 − f1p 1 − dm 1 − f1m 1 + log (cid:18) (cid:18) log (cid:19) (cid:32) dm f1m 1 + (cid:33) log dm f1m (cid:33) dp f1p (cid:18) (cid:19) dx β (cid:19) dx β DKL(P, P ) = fp(1 − f1p) 1 − dp + (1 − fp)(1 − f1m) + (1 − fp) f1m (cid:32) + fp f1p 1 + log 1 + . (10) This is a convex function of dx; we can see this by observing that each of the four terms is a function of the form F (dx) = α 1 + log 1 + (in the first term, for example, we have α = fp(1 − f1p) and β = −(1 − f1p)). This can be readily shown to be convex by taking the second derivative with respect to dx and verifying that it is positive: (cid:16) α 1 + dx β (cid:17), F (cid:48)(cid:48)(dx) = β2 23 where we can verify that α > 0 and dx < β. The sum of convex functions is likewise convex. Because DKL(P, Q) is non-negative for any distributions P and Q, DKL(P, P ) achieves its unique minimum along an iso-moment line at P = P , and it must monotonically increase as a function of dx. To get an intuitive picture of DKL(P, P ) as it varies in the (fp, f1p, f1m)-coordinate space, we view this quantity along constant-fp slices (Figure 2). DKL(P, P ) increases with distance from the constraint curve. Generally, the range of this distance peaks at fp = 0.25. To further quantify the relationship between distance and DKL, we approximate the logarithms in Equation 10 for small arguments and find that DKL increases quadratically with distance for small arguments: DKL(P, P ) ≈ (dx)2C(fp, f1m) + O(dx3) (11) where C(fp, f1m) = + fp(1 − fp)2/9 p + (1 − fp)2/9 f 2 p (1 − fp) f 2 p + (1 − fp)2/9 f 2 1m)2 (1 − 3 f1m + 3 f 2 1m(1 − f1m)3 f 3 1 f1m(1 − f1m) Numerical sampling of 3 cell network For general circuit set-ups, it may be necessary to probe the output distribution by sampling. In the case of global input, however, it is more computationally efficient and accurate to compute the output spiking probability distribution using quadrature. To be concrete, a set of N = 3 threshold spiking units is forced by a common input Ic (drawn from a probability distribution PC(y)) and an independent input Ij (drawn from a probability distribution PI(y)). The output of each cell xj is determined by summing and thresholding these inputs: xj = H(Ij + Ic − Θ) Conditioned on Ic, the probability of each spike is given by: Prob[xj = 1 Ic = a] = Prob[Ij + a − Θ > 0] = Prob[Ij > Θ − a] = PI(y) dy Similarly, we have the conditioned probability that xj = 0: Prob[xj = 0 Ic = a] = Prob[Ij + a − Θ < 0] = Prob[Ij < Θ − a] (cid:90) ∞ Θ−a (cid:90) Θ−a −∞ PI(y) dy = 24 Because these are conditionally independent, the probability of any spiking event (x1, x2, x3) = (A1, A2, A3) is given by the integral of the product of the conditioned probabilities against the density of the common input. Prob[x1 = A1, x2 = A2, x3 = A3] = dy PC(y) Prob[xj = Aj Ic = y] (12) (cid:90) ∞ −∞ 3(cid:89) j=1 The integrand in the previous equation is numerically evaluated via an adaptive quadrature routine, such at Matlab's quad. This is easily generalized to an arbitrary number of cells N . Unimodal inputs Ij, Ic were chosen from different marginals with mean 0 and variance σ2. For Gaussian input, P (x) ∝ e−x2/2σ2; for uniform inputs, P (x) ∝ 1 for x < 3σ2 and 0 otherwise. For skewed input, P (x) ∝ (x + µ)e−(x+µ)2/2a, for x > −µ, where the parameter a sets the variance 2a(1 − π 2 ensures that the mean of P (x) is zero. When sampling was necessary in thresholding models, for N ≤ 14 the number of samples was chosen so that the mean bin occupancy of each unique circuit output was > 100 (and usually > 1000). For N = 16 -- the most poorly resolved case -- the mean bin occupancy was lower ( 40). 4 ) and shifting by µ =(cid:112) aπ √ Bimodal triplet inputs always generate distributions with negative strain Another information theoretic quantity that relates to the ability of a distribution to be characterized by a pairwise model is the strain: (cid:18) P (1, 1, 1)P (1, 0, 0)P (0, 1, 0)P (0, 0, 1) (cid:19) γ = 1 8 log P (0, 0, 0)P (1, 1, 0)P (1, 0, 1)P (0, 1, 1) . Indeed, this quantity must be zero for any distribution that satisfies the PME constraint (Equation 2). Negative values of strain occur to the left side of the PME constraint curve in the (f1p, f1m)-plane, whereas positive values occur to the right. For a circuit forced by common binary inputs, the simplicity of our setup allows us to show why observed distributions occur to the left side of the PME constraint curve. We approach this by showing that given the f1m coordinate of an observed distribution, the f1p coordinate is less than the PME fit would predict. A point on the constraint surface corresponding to a particular value of f1m may be written f1p = f 3 1m 1 − 3f1m + 3f 2 1m whereas = q3 q3 + (1 − q)3 f1p = pq3 pq3 + (1 − p) + p(1 − q)3 25 and 1 f1p = = This makes it clear that 1 f1p ≥ 1 f1p 1 − p + p(q3 + (1 − q)3) 1 − p pq3 + pq3 1 f1p with equality if and only if p = 1. Therefore f1p < f1p (13) unless p = 1, in which case they coincide. According to Equation 13, p0 is under-predicted by the PME model, whereas p3 is over- predicted ; this is precisely the condition of "sparse coding" [11, 25]. An analytical explanation for unimodal vs. bimodal effects We consider an analytical argument to support the our numerical results that bimodal inputs generate larger deviations from PME model fits than unimodal inputs. As a metric, we consider DKL(P, P ) -- where P and P are again the true and model distributions respectively -- when we perturb an independent spiking distribution by adding a common, global input of variance c. To simplify notation, the small parameter in the calculation will be denoted  = √ We begin by observing that when P is a maximum entropy distribution; that is, its logarithm is given as a sum over functionals whose averages over P must also be satisfied by P , or c. log(p(x)) = p(x)fµ(x) = the KL-distance may be written as a difference of entropies: (cid:88) x p(x)fµ(x); (cid:33) µ x p(x) p((cid:126)x) log (cid:88) (cid:18) p(x) λµfµ(x) − log Z, (cid:88) (cid:19) DKL(P, P ) ≡ (cid:88) (cid:88) p(x) log(p(x)) −(cid:88) (cid:32) = −S(P ) −(cid:88) = −S(P ) + log Z −(cid:88) = −S(P ) + log Z −(cid:88) = −S(P ) −(cid:88) p(x) λµ λµ µ µ x x = x p(x) log(p(x)) p(x) log(p(x)) (cid:88) x − log Z + λµfµ(x) (cid:88) (cid:88) x x µ p(x)fµ(x) p(x)fµ(x) = −S(P ) + S( P ). x 26 Here, the entropy of a probability distribution P is given S(P ) = −p0 log(p0) − 3p1 log(p1) − 3p2 log(p2) − p3 log(p3) (14) if we use the fact that the distributions are permutation-symmetric (i.e. p1 ≡ P (1, 0, 0) = P (0, 1, 0) = P (0, 0, 1)). We take the logarithms in the definitions of entropy S and KL- divergence DKL to be base 2, so that any numerical values of these quantities are in units of bits. We now compute S(P ) and S( P ) by deriving a series expansion for each set of event prob- abilities. We can compute the true distribution P using the expressions derived in Equation 12; to recap, let the common input have probability density p(c), and the independent input to each cell have density ps(x). Let θ be the threshold for generating a spike (i.e., a "1" response). For each cell, a spike is generated if x + c > θ, i.e., with probability (cid:90) ∞ θ−c d(c) = ps(x)dx . Given c, this is conditionally independent for each cell. We can therefore write our proba- bilities by integrating over c as follows: −∞ (cid:90) ∞ (cid:90) ∞ (cid:90) ∞ (cid:90) ∞ −∞ −∞ −∞ p0 = p1 = p2 = p3 = p(c)(1 − d(c))3 dc p(c)d(c)(1 − d(c))2 dc p(c)d(c)2(1 − d(c)) dc p(c)d(c)3 dc (15) (cid:16)c (cid:17)  p(c) = 1  f We develop a perturbation argument in the limit of very weak common input. That is, p(c) is close to a delta function centered at c = 0. Take p(c) to be a scaled function (16) (cid:82) ∞ We place no constraints on f (c), other than that it must be normalized (E[1] = 1) and that its moments must be finite (so E[c], E[c2], and so forth must exist, where E[g(c)] ≡ −∞ g(c)f (c) dc). For the moment, assume that the function f (c) has a single maximum at c = 0. To evaluate the integrals above, we Taylor-expand d(c) around c = 0. Anticipating a sixth- order term to survive, we keep all terms up to this order. This gives, for small y, d(y) ≈ d(0) + akyk + O(y7) 6(cid:88) k=1 27 where a1 = ps(θ) (the other coefficients a2-a6 can be given similarly in terms of the indepen- dent input distribution at θ). Substituting this into the expressions for p0, etc., above, with p(c) given as in Equation 16, gives us each event as a series in ; for example, 0 +(cid:0)3a1d2 0 E[c](cid:1)  +(cid:0)(3a2 p3 = d3 0) E[c2](cid:1) 2 + ... 1d0 + 3a2d2 The entropy S(P ) is now given by using these series expansions in Equation 14. We note that our derivation does not rely on the fact that the distribution of common input is peaked at c = 0 in particular. For example, we could have a common input centered around µ. The common input distribution function would be of the form (cid:18) c − µ (cid:19)  p(c) = 1  f Changing  regulates the variance, but doesn't change the mean or the peak (assuming, without loss of generality, that the peak of f occurs at zero). The peak of p(c) now occurs at µ, and the appropriate Taylor expansion of d(y) is 6(cid:88) d(y) ≈ d(µ) + bk(y − µ)k + O(y7), k=1 where the coefficients bk now depend on the local behavior of d around µ. The expectations that appear in the expansion of p3, and so forth, are now centered moments taken around µ; the calculations are otherwise identical. In other words, perturbation expansion requires the variance of the common input to be small, but not the mean. For bimodal inputs, we consider a common input with a probability distribution of the following form: p(c) = (1 − 2) 1  f (cid:16)x (cid:17)  + 2 1  f (cid:18) x − 1 (cid:19)  so that most of the probability distribution is peaked at zero, but there is a second peak of higher order (here taken at c = 1, without loss of generality). Again, we approximate the integrals given in Equations 15, and therefore the entropy S(P ), by Taylor expanding d; 6(cid:88) 6(cid:88) k=1 d(c) ≈ d(0) + ≈ d(1) + akck + O(c7); (c ≈ 0) bk(c − 1)k + O((c − 1)7); (c ≈ 1) k=1 around the two peaks 0 and 1 respectively. For each integral we have the same contributions from the unimodal case, multiplied by (1 − 2), as well as the corresponding contributions from the second peak multiplied by 2 (these weightings are chosen so that the common 28 input has variance of order 2, as in the unimodal case). This makes clear at what order every term enters. We now construct an expansion for the PME model P : P (x1, x2, x3) = 1 Z exp (λ1(x1 + x2 + x3) + λ2(x1x2 + x2x3 + x1x3)) We approach this problem by describing λ1 and λ2 as a series in . We match coefficients by forcing the first and second moments of P to match those of P -- as they must. Specifically, take 6(cid:88) λ1 = λ + kuk + O(7) 6(cid:88) k=1 kvk + O(7) λ2 = where λ1 = λ, λ2 = 0 are the corresponding parameters from the independent case. The events p0, p1, p2 and p3 can be written as a series in . We then require that the mean and centered second moments of P match those of P ; that is k=1 p1 + 2p2 + p3 = p1 + 2p2 + p3 p2 + p3 − (p1 + 2p2 + p3)2 = p2 + p3 − (p1 + 2p2 + p3)2. At each order k, this yields a system of two linear equations in uk and vk; we solve, inductively, up to the desired order; we now have P , and therefore S( P ), as a series in . Finally, we combine the two series to find that in the unimodal case, DKL(P, P ) = S( P ) − S(P ) (cid:20)a6 (cid:21) = 6 1(2 E[c]3 − 3 E[c] E[c2] + E[c3])2 2(1 − d0)3d3 0 + O(7) (17) If the first two odd moments of the distribution are zero (something we can expect for "symmetric" distributions, such as a Gaussian), then this sixth-order term is zero as well. For the bimodal case DKL(P, P ) = S( P ) − S(P ) (cid:20) (d1 − d0)6 2(1 − d0)3d3 0 (cid:21) + O(5) = 4 This last term depends on the distance d1 − d0, in other words, how much more likely the independent input is to push the cell over threshold when common input is "ON". We can also view this as depending on the ratio d1−d0 , which gives the fraction of previously 1−d0 non-spiking cells that now spike as a result of the common input. The main point here, of course, is that DKL(P, P ) is of order 4 rather than 6. So, as the strength of a common binary vs. unimodal input increases, spiking distributions depart from the PME more rapidly. 29 Experimentally-based model of a RGC circuit We model the response of a individual RGC using data collected from a representative primate ON parasol cell, following methods in [32, 35]. Similar response properties were observed in recordings from 16 other cells. To measure the relationship between light stimuli and synaptic conductances, the retina was exposed to a full-field, white noise stimulus. The cell was voltage clamped at the excitatory (or inhibitory) reversal potential VE = 0 mV (VI = −60 mV), and the inhibitory (or excitatory) currents were measured in response to the stimulus. These currents were then turned into equivalent conductances by dividing by the driving force of ±60 mV; in other words I exc = gexc(V − VE); I inh = ginh(V − VI); V − VE = −60 mV V − VI = 60 mV The time-dependent conductances gexc and ginh were now injected into the same cell using a dynamic clamp (i.e., input current was varied rapidly to maintain the correct relationship between the conductance and the membrane voltage) and the voltage was measured at a resolution of 0.1 ms. To model the relationship between the light stimulus and synaptic conductances, the current measurements I exc and I inh were fit to a linear-nonlinear model: I exc(t) = N exc[Lexc ∗ s(t) + ηexc], I inh(t) = N inh[Linh ∗ s(t) + ηinh] where s is the stimulus, Lexc (Linh) is a linear filter, N exc (N inh) is a nonlinear function, and ηexc (ηinh) is a noise term. The linear filter was fit by the function Lexc(t) = Pexc(t/τexc)nexc exp(−t/τexc) sin(2πt/Texc) (18) and the nonlinear filter by the polynomial N exc = Aexcx2 + Bexcx + Cexc; k , ηinh Linh and N inh were fit using the same parametrization. The noise terms ηexc k were fit to reproduce the statistical characteristics of the residuals from this fitting. We simulated the noise terms ηexc and ηinh using Ornstein-Uhlenbeck processes with the appropriate parame- ters; these were entirely characterized by the mean, standard deviation, and time constant of autocorrelation τη,exc (τη,inh), as well as pairwise correlation coefficients for noise terms entering neighboring cells. The noise correlation coefficients were estimated from the dual recordings of [32]. Linear filter parameters computed were Pexc = −8 × 104 pA/s, nexc = 3.6, τexc = 12 ms, Texc = 105 ms, and Pinh = −1.8 × 105 pA/s, ninh = 3.0, τinh = 16 ms, Texc = 120 ms. Nonlinearity parameters were Aexc = −5 × 10−5, Bexc = 0.42, Cexc = −57, and Ainh = 1 × 104, Binh = 0.37, Cinh = 250. Noise parameters were measured to be mean(ηexc k ) = 30, 30 k ) = 500, τη,exc = 22 ms, and mean(ηinh std(ηexc addition, excitatory (inhibitory) noise to different cells ηexc coefficient of 0.3 (0.15). the filter function; i.e. Lexc,R(t) = Lexc(t). k ) = −1200, std(ηinh k , ηexc j k ) = 780, τη,inh = 33 ms. In (ηinh ) had a correlation k , ηinh j To obtain the rectified filter shown in Figure 7E, we simply took the absolute value of Model fitting We create a model of the cell as a nonlinear integrate-and-fire model using the method of Badel et al. [33], in which the membrane voltage is assumed to respond as dV dt = F (V, t − tlast) + Iinput(t) C (19) where C is the cell capacitance, tlast is the time of the last spike before time t, and Iinput(t) is a time-dependent input current. We use the current-clamp data, which yields cell voltage in response to the input current Iinput(t) = gexc(t)(V − VE) + ginh(V − VI), to fit a function F (V, t). When voltage data is segregated according to the (binned) time since the last spike, the I − V curve is well fit by a function of the form (cid:0)EL − V + ∆T e(V −VT )/∆T(cid:1) F (V ) = 1 τm (20) The membrane time constant τm, resting potential (EL), spike width ∆T and knee of the exponential curve VT are parameterized as a function of t−tlast. Our model neuron comprises Equations (19, 20) for V < Vthreshold; a spike was detected when V reached Vthreshold = −30 mV, with a voltage reset to Vreset = −55 mV after 2 ms. In addition, the cell was unable to spike for an absolute refractory period of τabs = 3 ms. The capacitance was inferred from the voltage trace data by finding, at a voltage value where the voltage/membrane current relationship is approximately Ohmic, the value of C that minimizes error in the relation Equation 19 [33]. The estimated value was C = 28 pF. Recurrent model Gap junction coupling was introduced as an additional current on the right-hand side of Equation 19: Igap,j C = −ggap C (Vj − Vk) (cid:88) k(cid:54)=j The coupling strength ggap was held constant during a simulation. When coupling was present (i.e. when ggap (cid:54)= 0), ggap was varied from a baseline level (1.1 nS) [32] to 16 times this value (17.6 nS) between simulations. For simulations that include electrotonic coupling, the spike trajectory was modeled with greater care. A typical (averaged) spike waveform was extracted from voltage traces of a 31 primate ON parasol cell. The spike waveform was used to replace 1 ms of the membrane voltage trajectory during and after a spike; at the end of the 1 ms, the voltage was released at approximately −58 mV. The cell was unable to spike for an absolute refractory period of τabs = 3 ms. A relative refractory period was induced by introducing a declining threshold for the period of 3-6 ms following a spike, after which Vth limits to -30 mV. Cell receptive field We defined each cell's stimulus as the linear convolution of an image with its receptive field. The receptive fields include an "on" center and an "off" surround, as in [40]: sj((cid:126)x) = exp ((cid:126)x − (cid:126)xj)Q((cid:126)x − (cid:126)xj) − k exp r((cid:126)x − (cid:126)xj)Qr((cid:126)x − (cid:126)xj) (cid:18) −1 2 (cid:19) (cid:18) −1 2 (cid:19) where the parameters k and 1/r give the relative strength and size of the surround. Q specifies the shape of the center and was chosen to have a 1 standard deviation (SD) radius of 50 µm and to be perfectly spherical. The receptive field locations (cid:126)x1, (cid:126)x2, and (cid:126)x3 were chosen so that the 1 SD outlines of the receptive field centers will tile the plane (i.e. they just touch). Other parameters used were k = 0.3, r = 0.675. Numerical methods and convergence testing For each simulation, Equations 19, 20 were integrated using the Euler method for > 105 ms with a time step of 0.1 ms. For each stimulus condition, 20 simulations (or sub-simulations) were run, for a total integration time of > 20 × 105 ms. The synaptic noise terms, ηexc and ηinh k , as well as the light input, were generated independently for each sub-simulation. To discretize spiking outputs, 5 ms bins were used. k These 20 sub-simulations were used to estimate standard errors in both the probabil- ity distribution over spiking events and DKL(P, P ). For example, in the constant light case, we generated the following distribution on spiking events: P (0, 0, 0) = 0.816 ± 0.004, P (0, 0, 1) = 0.0457 ± 0.0015; P (0, 1, 0) = 0.0448 ± 0.0015, P (1, 0, 0) = 0.0459 ± 0.0017, P (0, 1, 1) = 0.00554± 0.00054, P (1, 0, 1) = 0.00545± 0.00051, P (1, 1, 0) = 0.00545± 0.00056, and P (1, 1, 1) = 0.00116 ± 0.00020. Numbers reported in the Results are, unless specified otherwise, produced by collating the data from the 20 simulations. To test our finding that the observed distributions were well-modeled by the PME fit, we also performed the PME analysis on each of the 20 simulations for each stimulus condition. While in general DKL(P, P ) can be quite sensitive to perturbations in P , the numbers re- mained small under this analysis. To confirm that our results for DKL(P, P ) are sufficiently resolved to remove bias from sampling, we performed an analysis in which we collect the 20 simulations in subgroups of 1, 2, 4, 5, 10, and 20, and plot the mean DKL with estimated standard errors. As expected (e.g. [41]), bias decreases as the length of subgroup increases and asymptotes at -- or before -- the full simulation length. Results are shown in Figure S1 for the RGC simulations under full-field stimulation, as well as two representative cases with "stixel" stimuli. 32 For the thresholding model, sampling was used for pairwise input cases in Figure 11, and all data presented in Figure 12. In all but three simulations, the sampling frequency was chosen so that the mean bin occupancy (that is, the average frequency of each unique spiking pattern) was > 100 (and usually > 1000). The exceptions are for the cases where the network size N = 16 and unimodal pairwise inputs are used, for which the mean bin occupancy was ≈ 40. We show the bias for N = 16, pairwise Gaussian inputs in Figure S1A, which confirms that the bias is close to its asymptotic value of zero. To provide a cross-validation test, we divided our data into halves (which we denote P1 and P2, each including data from 10 subsimulations) and performed the PME analysis on one half (say P1) to yield a model P1. We then computed DKL(P2, P1) and DKL(P2, P1) (as in [42]), which we refer to the cross-validated and empirical likelihood respectively. The former tests whether the PME fit is robust to over-fitting; the latter tests how well-resolved our "true" distribution is in the first place. In Figure S2, we plot these numbers vs. the numbers reported in the paper. Most cross-validated likelihoods fall on or near the identity line; most empirical likelihoods are close to zero (and importantly, significantly smaller than either DKL(P, P ) or DKL(P2, P1), indicating that DKL(P, P ) is accurately resolved). We conclude that the deviations that we observe when these conditions are met can not be accounted for by the differences in testing and training data. Finally, to test the robustness of our finding that the strain γ < 0 for the full-field RGC simulations, we perturbed our spiking event distributions randomly, with perturbations weighted by the estimated standard errors. This was repeated 20 times for each stimulus condition. Out of the resulting 840 perturbations (20 each for 42 stimulus conditions), γ > 0 in only 22 trials. Financial Disclosure This research was supported by NSF grant DMS-0817649 and by a Career Award at the Scientific Interface from the Burroughs-Wellcome Fund (ESB), by the Howard Hughes Med- ical Institute and by NIH grant EY-11850 (FMR), and a Cambridge Overseas Research Studentship (JG). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. 33 References [1] Baudry M, Taketani M, editors (2006) Advances in Network Electrophysiology Using Multi-Electrode Arrays. New York: Springer Press. [2] Amari S (2001) Information geometry on hierarchy of probability distributions. IEEE Transactions on Information Theory . [3] Martignon L, Deco G, Laskey K, Diamond M, Freiwald W, et al. (2000) Neural coding: higher-order temporal patterns in the neurostatistics of cell assemblies. Neural Comp 12: 2621 -- 53. [4] Schneidman E, Still S, Berry (II) MJ, Bialek W (2003) Network information and con- nected correlations. Physical Review Letters 91: 238701. [5] Shlens J, Field GD, Gauthier JL, Grivich MI, Petrusca D, et al. (2006) The structure of multi-neuron firing patterns in primate retina. Journal of Neuroscience 26: 8254 -- 8266. [6] Shlens J, Field GD, Gauthier JL, Greschner M, Sher A, et al. (2009) The structure of large-scale synchronized firing in primate retina. Journal of Neuroscience 29: 5022 -- 5031. [7] Schneidman E, Berry (II) MJ, Segev R, Bialek W (2006) Weak pairwise correlations imply strongly correlated network states in a neural population. Nature 440: 1007 -- 1012. [8] Tang A, Jackson D, Hobbs J, Smith JL, Patel H, et al. (2008) A maximum entropy model applied to spatial and temporal correlations from cortical networks in vitro. Journal of Neuroscience 28: 505 -- 518. [9] Yu S, Huang D, Singer W, Nikolic D (2008) A small world of neuronal synchrony. Cerebral Cortex 18: 2891 -- 2901. [10] Montani F, Ince RAA, Senatore R, Arabzadeh E, Diamond ME, et al. (2009) The impact of high-order interactions on the rate of synchronous discharge and information transmission in somatosensory cortex. Philosophical Transactions of the Royal Society A 367: 3297 -- 3310. [11] Ohiorhenuan IE, Mechler F, Purpura KP, Schmid AM, Hu Q, et al. (2010) Sparse coding and high-order correlations in fine-scale cortical networks. Nature 466: 617 -- 21. [12] Santos GS, Gireesh ED, Plenz D, Nakahara H (2010) Hierarchical interaction structure of neural activities in cortical slice cultures. The Journal of Neuroscience 30: 8720 -- 8733. [13] Ganmor E, Segev R, Schneidman E (2011) Sparse low-order interaction network un- derlies a highly correlated and learnable population code. Proceedings of the National Academy of Sciences USA 108: 9679 -- 9684. 34 [14] Amari S, Nakahara H, Wu S, Sakai Y (2003) Synchronous firing and higher-order inter- actions in neuron pool. Neural Computation 15: 127 -- 142. [15] Krumin M, Shoham S (2009) Generation of spike trains with controlled auto- and cross- correlation functions. Neural Comp 21: 1642 -- 64. [16] Macke JH, Berens P, Ecker AS, Tolias AS, Bethge M (2009) Generating spike trains with specified correlation coefficients. Neural Computation 21: 397 -- 423. [17] Roudi Y, Nirenberg S, Latham PE (2009) Pairwise maximum entropy models for study- ing large biological systems: When they can work and when they cant. PLoS Compu- tational Biology 5: e1000380. [18] Roudi Y, Tyrcha J, Hertz J (2009) Ising model for neural data: Model quality and approximate methods for extracting functional connectivity. Physical Review E 79: 051915. [19] Bohte SM, Spekreijse H, Roelfsema PR (2000) The effects of pair-wise and higher-order correlations on the firing rate of a postsynaptic neuron. Neural Computation 12: 153 -- 179. [20] Macke JH, Opper M, Bethge M (2011) Common input explains higher-order correlations and entropy in a simple model of neural population activity. Physical Review Letters 106: 208102. [21] Cocco S, Leibler S, Monasson R (2009) Neuronal couplings between retinal ganglion cells inferred by efficient inverse statistical physics methods. Proceedings of the National Academy of Sciences USA 106: 14058 -- 14062. [22] Tkacik G, Schneidman E, Berry II MJ, Bialek W (2009) Spin glass models for a network of real neurons. ArXiv:0912.5409. [23] Jaynes ET (1957) Information theory and statistical mechanics. Physiol Rev 106: 620- 630. [24] Jaynes ET (1957) Information theory and statistical mechanics II. Physiol Rev 108: 171-190. [25] Ohiorhenuan IE, Victor JD (2010) Information-geometric measure of 3-neuron firing patterns characterizes scale-dependence in cortical networks. Journal of Computational Neuroscience : 10.1007/s10827-010-0257-0. [26] Cover TM, Thomas JA (1991) Elements of information theory. New York: Wiley. [27] de la Rocha J, Doiron B, Shea-Brown E, Josic K, Reyes A (2007) Correlation between neural spike trains increases with firing rate. Nature 448: 802 -- 806. 35 [28] Vilela RD, Lindner B (2009) Comparative study of different integrate-and-fire neurons: Spontaneous activity, dynamical response, and stimulus-induced correlation. Physical Review E 80: 031909. [29] Shea-Brown E, Josi´c K, Doiron B, de la Rocha J (2008) Correlation and synchrony transfer in integrate-and-fire neurons: Basic properties and consequences for coding. Phys Rev Lett 100: 108102. [30] Marella S, Ermentrout GB (2008) Class-II neurons display a higher degree of stochastic synchronization than class-I neurons. Physical Review E 77: 041908. [31] Barreiro AK, Shea-Brown ET, Thilo EL (2010) Timescales of spike-train correlation for neural oscillators with common drive. Physical Review E 81: 011916. [32] Trong PK, Rieke F (2008) Origin of correlated activity between parasol retinal ganglion cells. Nature Neuroscience 11: 1343 -- 1351. [33] Badel L, Lefort S, Brette R, Petersen CCH, Gerstner W, et al. (2007) Dynamic I-V curves are reliable predictors of naturalistic pyramidal-neuron voltage traces. Journal of Neurophysiology 99: 656 -- 666. [34] Sharpe LT, Whittle P, Nordby K (1993) Spatial integration and sensitivity changes in the human rod visual system. J Physiol 461: 235-246. [35] Murphy GJ, Rieke F (2006) Network variability limits stimulus-evoked spike timing precision in retinal ganglion cells. Neuron 52: 511-524. [36] Dacey D, Brace S (1992) A coupled network for parasol but not midget ganglion cells in the primate retina. Vis Neurosci 9: 279-290. [37] Bethge M, Berens P (2008) Near-maximum entropy models for binary neural represen- tations of natural images. In: Advances in Neural Information Processing Systems 20. pp. 97 -- 104. [38] Oizumi M, Ishii T, Ishibashi K, Okada M (2010) Mismatched decoding in the brain. Journal of Neuroscience 30: 4815 -- 4826. [39] Dayan P, Abbot L (2001) Theoretical Neuroscience: Computational and Mathematical Modeling of Neural Systems. MIT Press, 1st edition. [40] Chichilnisky EJ, Kalmar RS (2002) Functional asymmetries in ON and OFF ganglion cells of primate retina. Journal of Neuroscience 22: 2737 -- 2747. [41] Paninski L (2003) Estimation of entropy and mutual information. Neural Computation 15: 1191 -- 1253. [42] Yu S, Yang H, Nakahara H, Santos G, Nikoli´c D, et al. (2011) Higher-order interactions characterized in cortical activity. Journal of Neuroscience 31: 17514 -- 17526. 36 Figure 1: Schematic showing different axes on which we explore microcircuits in this study. (Top left) Network architecture: global vs. pairwise inputs and scaling up system size N . (Top right) Presence vs. absence of reciprocal coupling. (Bottom left) Input statistics: unimodal vs. bimodal marginal statistics. (Bottom right) varying strength of common input. 37 vsvsvsIcI1I2I3I13I12I23I1I2IN-1INI1I12I23IN-1,NIN,1I2IN1 0 0 1 0 ...0 1 0 1 0 ...0 1 0 0 1 ...IcIcI1I2I3vsIcI1I2I3 Figure 2: Geometrical organization of DKL(P, P ) within the space of three-cell spiking distributions. Outer plots: Slices of DKL(P, P ) along surfaces fp = constant. Cen- ter: schematic of the (fp, f1p, f1m) coordinate space. Counterclockwise from lower right: fp = 0.1, 0.25, 0.58, 0.76, 0.9. fp = 0.25 contains the maximal attainable DKL(P, P ) over all admissible P . fp = 0.616 contains the maximal attainable DKL from pairwise bimodal inputs (see Results). 38 fpf1pf1mf1mf1pfp00.5100.20.40.60.81f1mf1pfp00.5100.20.40.60.81= 0.76f1mf1pfp00.5100.20.40.60.81= 0.58f1mf1pfp00.5100.20.40.60.81= 0.25f1mf1pfp00.5100.20.40.60.81= 0.1= 0.9 Figure 3: Examples of spiking distributions with small, intermediate, and large deviations from PME models (as quantified by the KL- divergence, DKL(P, P ), between the observed distribution P and its PME fit P ). (Top row) Bar plot contrasting three distributions with their pairwise maximum entropy (PME) fits. The probability shown is the total probability of all events with a certain number of spikes. From left: global skewed input, global binary input, XOR operator. DKL(P, P ) is, from left, 0.0013 (skewed), 0.091 (bimodal), and 1 (XOR). (Middle row) The same distributions (crosses) projected into the (f1m, f1p)-plane and their corresponding PME fits (circles). The cyan line is the iso-moment line of all distributions with the same first and second moments, and the black curve is the PME constraint surface. (Bottom row) DKL(P, P ) along the iso-moment line (cyan solid) and the quadratic approximation derived in the Methods. (black dashed). 39 00.20.40.60.8100.20.40.60.8100.20.40.60.8100.20.40.60.81012300.20.40.60.81# spikes# spikes# spikesP(s)f1mf1mf1mf1pf1mf1mf1mbitsglobal skewed inputglobal bimodal inputXOR operationObservedPMEObservedPMEObservedPMED = 0.0013D = 0.091D = 1KLKLKL Figure 4: Deviation from PME fit for circuits receiving independent and (global) common input. Results shown for N = 3; (A) Gaussian, (B) uniform, (C) skewed, and (D) bimodal. For each choice of marginal input statistics, possible input parameters are varied over a broad range as described in the Results; over c ∈ [0, 1], σ ∈ [0, 4], and Θ ∈ [−1, 3] (unimodal inputs), over p ∈ [0, 1] and q ∈ [0, 1] (bimodal inputs). (Left column) Schematic of input dis- tributions. (Center column) Projection of all distributions onto the (f1p, f1m)-plane. (Right column) For the value of Θ for which the maximal value is achieved, a slice of DKL(P, P ) (unimodal inputs); contour plot of DKL(P, P ) where 1 < Θ < 2 (bimodal inputs). 40 qABCD Figure 5: Relationship between measures of higher-order interactions and other output firing statistics. (A-C) DKL(P, P ) versus firing rate E[xj], for data obtained from N = 3 threshold- ing cells. For each choice of marginal input statistics, possible input parameters were varied over a broad range as described in the Results; over c ∈ [0, 1], σ ∈ [0, 4], and Θ ∈ [−1, 3] for unimodal inputs, and over p ∈ [0, 1] and q ∈ [0, 1] for bimodal inputs. In each panel, data is organized by ρ, from dark (ρ ∈ (0, 0.1)) to light (ρ ∈ (0.9, 1)); (A) Gaussian inputs, (B) skewed inputs, (C) bimodal inputs. (D-F) The fraction of multi-information captured by the PME model, ∆, versus firing rate E[x1], for these same distributions. Data is organized by correlation coefficient ρ, as in panels (A-C); D) Gaussian, E) skewed, F) bimodal. 41 Figure 6: Effect of recurrent coupling on PME fits of circuit outputs for the thresholding model. Left panel: scatter plot of DKL achieved with Gaussian inputs vs. coupling strength g, for N = 3 cells. Right panel: another view of DKL(P, P ) vs. coupling strength g, for N = 3 cells. In the top three panels, for unimodal inputs, c, σ, and Θ were fixed at a single representative value, while g was varied. The top panel shows the results from circuits with Gaussian inputs (global and pairwise inputs are equivalent) and the next two for skewed and uniform inputs with either global (thin lines) or local (heavy lines) inputs. For all unimodal inputs, (c, σ, Θ) were (0.5, 1, 1). Analogous results from circuits with either global (thin line) or local (heavy line) bimodal inputs are shown in the bottom panel. Here, p, q, and Θ were fixed at a single representative value, while g was varied. The parameters (p, q, Θ) were (0.81, 0.5, 1.5) and (0.5, 0, 1.5) for global and pairwise inputs respectively. 42 00.511.522.530123456789x 10-3g (coupling strength)D (P, P)KL~A00.511.5200.0050.0100.010.020.0500.511.50.20.40000.511.5200.511.52g (coupling strength)D (P, P)~KLB 43 -1500-500050010002468101214-1000x 104-2000-1000010000.511.522.533.544.5x 104-2000-10000100002468101214x 104-3000-2000-100001000024681012141618x 104ABCD00.20.40.60.8ObservedPME-1-0.500.51050100150200−1001000Lexc050100150200−1000100Lexc0400800120016002000-1-0.500.51E−3000−2000−10000100000.511.52x 10500.20.40.60.8000100110111ObservedPME00010011011100010011011100.20.40.60.8ObservedPME00010011011100010011011100.20.40.60.8ObservedPME00.20.40.60.8ObservedPME−10400800120016002000-2-1.5-1-0.500.511.520.511.52502004006008001000-1-0.500.512.5x 10time (ms)% dev. from meantime (ms)-2-1.5-1-0.500.511.52012x 1002004006008001000-1-0.500.5143% dev. from meantime (ms)-2-1.5-1-0.500.511.525101502004006008001000-1-0.500.51x 104% dev. from meantime (ms)time (ms)% dev.% dev.% dev.countcountcount% dev./pA% dev./pA1000-20001000-2000pAcountpAcountpAcountpAcountpAcount Figure 7: Composite of results for RGC simulations with constant light and full field flicker. In rows (A-C), we have (left) a histogram and time series of stimulus, (center) a histogram of excitatory conductances (multiplied by driving voltage −60 mV, and therefore in units of pA) and (right) the resulting distribution on spiking patterns. (A) Gaussian noise only. (B) Gaussian input, standard deviation 1/6, refresh rate 8 ms. (C) Binary input, standard deviation 1/3, refresh rate 8 ms. (D-E) Binary input, standard deviation 1/3, refresh rate 100 ms. In the top left panels, the excitatory filter Lexc(t) is shown instead of a stimulus histogram; in the bottom left panels, the normalized excitatory conductance (in pA -- red dashed line) is superimposed on the stimulus (blue solid). (D) Filter fit from data, parameters given in Methods. Both the filter and conductance trace illustrate that the LN model that processes light input acts as a (time-shifted) high pass filter. (E) As in (D), but with rectified filter, producing a bimodal distribution of conductances. (A) DKL(P, P ) for original vs. simulations incorporating a rectified filter. Figure 8: Comparison of RGC simulations computed with the original ON parasol filter, vs. rectified filter. Data is organized by stimulus refresh rate (8, 40, and 100 ms) and marginal statistics (Gaussian vs. binary). (B-C) Histograms of excitatory conductances (multiplied by driving voltage −60 mV, and therefore in units of pA) for representative simulations, under original (left) and rectified (right) filters. (B) Binary stimulus with standard deviation of 1/4 and refresh rate 100 ms. The histogram shows a significant shift towards a bimodal structure, with a corresponding increase in DKL(P, P ). (C) Binary stimulus with standard deviation of 1/2 and refresh rate 8 ms. The histogram shows little qualitative change under the change in filter, and little change in DKL(P, P ). 44 00.0020.0040.0060.0080.010.01200.0020.0040.0060.0080.010.0128 ms40 ms100 msGaussianBinaryD (P, P), original filterKL~D (P, P), rectified filterKL~ABCcountpAcountpApApA 45 02004000246x 10−3A00.0500.08stixel size = 256 mµP(1,0,0), P(1,1,0)P(0,0,1), P(0,1,1)0.060.040.020.040.080P(0,1,0),P(1,0,1)B0.0500.040.08stixel size = 60 mµP(1,0,0), P(1,1,0)P(0,1,0),P(1,0,1)0.060.02000.040.08P(0,0,1), P(0,1,1)C321D−2002040600510x 105−2002040600510x 105−2002040600510x 105132E00.20.40.6p0p1p2p3ObservedPMEF321G−2002040600510x 105−2002040600510x 105−2002040600510x 105123H00.20.40.6p0p1p2p3ObservedPMEI Figure 9: Results for RGC simulations with light stimuli of varying spatial scale ("stixels"). With the exception of (A), we show data from the binary light distributions; results from the Gaussian case are similar. (A) Average DKL(P, P ) as a function of stixel size. Values were averaged over 5 stimulus positions, each with a different (random) stimulus rotation and translation; 512 µm corresponds to full field stimuli. (B,C) Probability of singlet and doublet spiking events, under stimulation by movies of 256 µm (B) and 60 µm (C) stixels. Event probabilities are plotted in 3-space, with the x, y, and z axes identifying the singlet (doublet) events 001 (011), 010 (101), and 100 (110) respectively. The black dashed line indicates perfect cell-to-cell homogeneity (e.g. P (1, 0, 0) = P (0, 1, 0) = P (0, 0, 1)). Both individual runs (dots) and averages over 20 runs (large circles) are shown, with averages outlined in black (singlet) and gray (doublet). Different colors indicate different stimulus positions. (D-F) Results for one stimulus position, with stixel size 60 µm. (D) Contour lines of the three receptive fields (at 0.5, 1, 1.5 and 2 SD; and at the zero contour line) superim- posed on the stimulus checkerboard (for illustration, pictured in an alternating black/white pattern). (E) Marginal distributions of the excitatory conductances, for each cell. (F) Spike pattern distribution; the three different probabilities labeled p1 correspond to, e.g, P (1, 0, 0), P (0, 1, 0), and P (0, 0, 1)), demonstrating heterogenous responses among the RGCs. (G-I) As in (D-F), but for stixel size 256 µm. 46 Figure 10: The impact of recurrent coupling on RGC networks with full-field visual stimuli. The strength of gap junction connections was varied from a baseline level (relative magnitude g = 1, or absolute magnitude ggap = 1.1 nS) to an order of magnitude larger (g = 16, or ggap = 17.6 nS). (A) DKL(P, P ) obtained from two independent simulations without coupling, for each of 42 different stimulus ensembles (deviation of this data from the line y = x thus gives a control for the statistical variability of the results in later panels). (B-F) DKL(P, P ) obtained with coupling, plotted versus the value obtained for the same stimulus ensemble without coupling, for each of 42 different stimulus ensembles. 47 000.0050.010.0050.01ggap= 0 nSggap= 1.1 nSggap= 2.2 nSggap= 4.4 nSggap= 8.8 nSggapDKL, no coupling= 17.6 nS00.0050.010.0050.01000.0050.010.0050.01000.0050.010.0050.01000.0050.010.0050.01000.0050.010.0050.01DKL, with couplingDKL, with couplingDKL, no couplingDKL, no coupling8 msGaussianBinaryABCFED Figure 11: Maximal deviation from PME fit for thresholding circuit model, forced by either global or local input against background noise, for increasing network size N . For each N , possible input parameters are ranged as described in the Results: over c ∈ [0, 1], σ ∈ [0, 4], and Θ ∈ [−1, 3] (unimodal inputs), over s ∈ [0, 1] and p ∈ [0, 1] (bimodal inputs). The sidebar shows sample distributions with maximal DKL(P, P ) for N = 16; from top, global bimodal inputs, pairwise bimodal inputs, global Gaussian inputs, and pairwise Gaussian inputs. 48 4681012141600.020.040.060.080.10.120.140.16pairwise, bimodalpairwise, uniformpairwise, skewedpairwise, gaussianglobal, bimodalglobal, uniformglobal, skewedglobal, gaussianbits per neuronN04812160.20.4004121600.20.48ObservedPME048121600.10.2048121600..13 Figure 12: Maximal deviation from PME fit for circuit forced by either global or local input against background noise and either global or local recurrent coupling. Plot shows the maximal DKL achieved over all parameters as N increases. For each N , possible input parameters are ranged described as in the Results; over c ∈ [0, 1], σ ∈ [0, 4], and Θ ∈ [−1, 3] (unimodal inputs), over s ∈ [0, 1], p ∈ [0, 1], and Θ ∈ [0.3, 1.9] (bimodal inputs). In addition, recurrent coupling strength g is varied over [0, 2.4]. The sidebar shows sample distributions with maximal DKL(P, P ) for N = 8; from top, global bimodal inputs with global coupling; pairwise bimodal inputs with global coupling; global Gaussian inputs with pairwise coupling. 49 Nbits per neuronglobal input, global couplingglobal input, local couplinglocal input, global couplinglocal input, local couplingbimodaluniformskewedgaussian00.10.30.50.7246802468000.10.30.52468000.20.40.60.8ObservedPME Figure 13: The fraction of multi-information captured by the PME model, ∆, is low for some network parameters as population size increases. Here, we show data obtained from feedforward networks of N = 12 thresholding cells. For each choice of marginal input statistics, possible input parameters were varied over correlation coefficient c ∈ [0, 1] and input variance σ ∈ [0, 4]. The threshold was held at Θ = 1. (A-B) ∆ for (A) Gaussian and (B) skewed input, for N = 12. (C) 1 − ∆ vs. N for Gaussian, skewed and uniform inputs at a fixed value of c and σ (c = 0.56; for Gaussian and skewed, σ = 0.2; for uniform, σ = 0.6). Because of the low firing rate, 1 − ∆ must grow linearly with N (see text). 50 Figure S1: Mean and estimated standard errors of DKL(P, P ), as a function of subgroup size. The 20 simulations for each circuit condition (see Materials and Methods) were collected into subgroups of M = 1, 2, 4, 5, 10, and 20. M = 20 corresponds to the full simulation length -- 2× 106 ms in (B), 2× 105 ms in (C,D) -- reported in the text. As expected, bias decreases as the length of subgroup increases and asymptotes at -- or before -- the full simulation length. (A) N = 16, Gaussian pairwise inputs, for the sum-and-threshold model. (B) Full-field RGC simulations. (C) Spatially variable RGC simulations, binary stimulus, stixel size = 4 µm. Different colors signify different positions of stimulus relative to receptive field. (D) As in (C), but stixel size = 256 µm. 51 0510152000.40.81.21.62x 10−4M (# simulations per subgroup)M (# simulations per subgroup)M (# simulations per subgroup)M (# simulations per subgroup)D (P, P)KL~D (P, P)KL~D (P, P)KL~D (P, P)KL~ABCD Figure S2: Cross-validated and empirical values of DKL(P, P ), vs. values reported in the paper. (A) Full field, (B) Rectified, (C) Stixel simulations with binary inputs, (D) Same as (C), zoomed into origin. 52 Cross−validated8 ms40 ms100 msGaussianBinaryACross−validated8 ms40 ms100 msGaussianBinaryB0246x 10−301234567x 10−3Cross−validatedEmpirical model4-32 x 10 m60 x 10 m128 x 10 m256 x 10 m-6-6-6-6CD4-32 x 10 m60 x 10 m128 x 10 m256 x 10 m-6-6-6-6
1611.05918
1
1611
2016-11-17T22:08:20
Morphology of Fly Larval Class IV Dendrites Accords with a Random Branching and Contact Based Branch Deletion Model
[ "q-bio.NC" ]
Dendrites are branched neuronal processes that receive input signals from other neurons or the outside world [1]. To maintain connectivity as the organism grows, dendrites must also continue to grow. For example, the dendrites in the peripheral nervous system continue to grow and branch to maintain proper coverage of their receptor fields [2, 3, 4, 5]. One such neuron is the Drosophila melanogaster class IV dendritic arborization neuron [6]. The dendritic arbors of these neurons tile the larval surface [7], where they detect localized noxious stimuli, such as jabs from parasitic wasps [8]. In the present study, we used a novel measure, the hitting probability, to show that the class IV neuron forms a tight mesh that covers the larval surface. Furthermore, we found that the mesh size remains largely unchanged during the larval stages, despite a dramatic increase in overall size of the neuron and the larva. We also found that the class IV dendrites are dense (assayed with the fractal dimension) and uniform (assayed with the lacunarity) throughout the larval stages. To understand how the class IV neuron maintains its morphology during larval development, we constructed a mathematical model based on random branching and self-avoidance. We found that if the branching rate is uniform in space and time and that if all contacting branches are deleted, we can reproduce the branch length distribution, mesh size and density of the class IV dendrites throughout the larval stages. Thus, a simple set of statistical rules can generate and maintain a complex branching morphology during growth.
q-bio.NC
q-bio
Morphology of Fly Larval Class IV Dendrites Accords with a Random Branching and Contact Based Branch Deletion Model. Sujoy Ganguly1, Olivier Trottier1, Xin Liang2, Hugo Bowne-Anderson1, and Jonathon Howard1 1Department of Molecular Biophysics and Biochemistry, Yale University, New Haven 2Tsinghua-Peking Joint Center for Life Sciences, School of Life Sciences, Tsinghua University, Beijing, China April 16, 2018 Abstract Dendrites are branched neuronal processes that receive input signals from other neurons or the outside world [1]. To maintain connectivity as the organism grows, dendrites must also continue to grow. For example, the dendrites in the peripheral nervous system continue to grow and branch to maintain proper coverage of their receptor fields [2, 3, 4, 5]. One such neuron is the Drosophila melanogaster class IV dendritic arborization neuron [6]. The dendritic arbors of these neurons tile the larval surface [7], where they detect localized noxious stimuli, such as jabs from parasitic wasps [8]. In the present study, we used a novel measure, the hitting probability, to show that the class IV neuron forms a tight mesh that covers the larval surface. Furthermore, we found that the mesh size remains largely unchanged during the larval stages, despite a dramatic increase in overall size of the neuron and the larva. We also found that the class IV dendrites are dense (assayed with the fractal dimension) and uniform (assayed with the lacunarity) throughout the larval stages. To understand how the class IV neuron maintains its morphology during larval development, we constructed a mathematical model based on random branching and self-avoidance. We found that if the branching rate is uniform in space and time and that if all contacting branches are deleted, we can reproduce the branch length distribution, mesh size and density of the class IV dendrites throughout the larval stages. Thus, a simple set of statistical rules can generate and maintain a complex branching morphology during growth. In our brains, billions of neurons interact with each other to build a nervous system of unparalleled complexity and computational power. Neurons have dendrites, which are branched structures that receive synaptic or sensory inputs, and an axon, which send outputs to other neurons. The shape or morphology of individual neurons sets the number and types of interactions that a neuron can have and provides the structural basis of neuronal computation [9, 10, 11, 12, 13, 14]. Since many organisms continue to enlarge after the establishment of the body plan, it is critical for axons and dendrites to maintain their morphology as they grow. For example, interneurons of the grasshoper [2], motor neurons in moths [3] and mice [4] grow drastically in size yet maintain connections to their target cells. Futhermore, dendrites in the perpherial nervous system, like those of gold fish retinal ganglion cells [5], and dendritic arborization (da) sensory neurons of the fly larva [6], which are the subject of this work, grow to continually maintain coverage of their receptor fields. In this paper, we investigate the growth rules that are required to maintain the correct branching morphology as a dendrite grows. The da sensory neurons of the fly larva are a model system for studying dendritic arborization [15, 16, 17]. These dendrites innervate the extracellular matrix, which lies between the outer cuticle and the inner epidermal cell layer [7]. They tile the surface on the fly larva in a highly stereotyped manner and have four distinct morphological classes [18] (Fig. 1 A). Since it is easy to identify and image individual da neurons, these neurons have proven to be a powerful model system for studying dendrite morphology [15, 16]. In this paper, we address the question of how the morphology of the class IV da neurons (Fig. 1 B) is maintained during the larval stages. The class IV da neuron has highly branched dendrites [18], which detect potentially harmful stimuli, such as the ovipositor barb of parasitic wasps [8, 19]. The dendrites of the class IV neuron begin 1 morphogenesis during late embryogenesis ∼ 16 hrs After Egg Lay (AEL). By the time the larva hatches (∼ 22 hrs AEL at 25 ◦C), the class IV dendrites nearly cover its surface. The dendrites then continue to expand and branch as the larva grows (22 − 126 hrs AEL), so that the neuron maintains its coverage of the larval surface [6]. In this work, we are seeking the growth rules that allow class IV dendrites to maintain their dense coverage of the larval surface. To this end, we have used a novel measure, the hitting probability, that quantifies the mesh size and two well-known measures of branching morphology: the fractal dimension [20] and lacunarity [21, 22] (see Definition of Morphometrics). We show that these measures remain largely invariant over larval stages, despite a several fold increase in larval length. Furthermore, we demonstrate that a model with simple rules for branching and self-avoidance can capture essential features of the establishment and maintenance of the dendrite's morphology. 1 Experimental Results To characterize the morphology of fully-developed class IV dendrites, we imaged larvae expressing Cd4- tdGFP under the ppk promoter (ppk-cd4-tdGFP ) during the third instar stage (Fig. 1) using a laser- scanning confocal microscope (See Material and Methods for details). Using NeuronStudio [23] and Fiji we traced the branches of the dendrites to produce skeletons. These skeletons were then analyzed to obtain the the mesh size, density and uniformity of class IV dendrites using parameters defined in the next section. Definition of Morphometrics Here we include simple definitions of the relevant morphometrics to aid comprehension. Hitting Probabiltiy H(B): The probability that a box of size B hits the dendrite. Mesh Size BH: The length at which 50% of all boxes hit the neuron. Fractal Dimension df: A measure of the space-fillingness of a shape. For a completely filled box df = 2, for a straight line df = 1, for branched shapes 1 ≤ df ≤ 2. Lacunarity Λ(B): A measure of density fluctuations as a function of length scale B. Lacunarity Length BΛ: The length at which Λ(B = BΛ) = 0.25, i.e. the length at which the neuron is uniform. The larger BΛ, the more variable the density of the neuron. Radius of Gyration Rg: A length scale that measures how spread out a shape is from its center. The larger Rg the more spread out the neuron. Persistence Length β: The characteristic length at which a branch bends. For mathematical definitions see Appendix. 1.1 Class IV dendrites have a small mesh size To characterize the mesh size of the dendrites, we developed a novel measure called the hitting probability H(B). H(B) measures the probability H that a randomly placed box of size B hits the dendrite (see Appendix for details). The hitting probability generalizes an earlier metric called the coverage index [6] by allowing for any box location and any box size. A typical hitting probability curve of a neuron (Fig. 1 D) H(B) increases monotonically with B, eventually reaching H = 1 as B approaches the size of the neuron. We define the characteristic mesh size BH as the box size at which half of all boxes hit the dendrite. In other words, BH is the maximum size of a stimulus that would go untouched, or undetected, on average, by the neuron. BH is similar to the mesh size in a cross-linked polymer network[24]. We found that BH = 8.4 ± 0.5 µm (mean ± SD, n = 14 neurons) for the mature dendrites of the class IV neuron. Thus, the mesh size is approximately equal to the diameter of the ovipositor barb of wasps that lay eggs in Drosophila larva (∼ 10 µm, [8]). This indicates that the class IV dendrite has a high chance of detecting a wasp attack. Furthermore, the mesh size is small compared to the overall size of the neuron (∼ 500 µm) and is similar to the mean branch length (see below). 2 Figure 1: Morphometrics of class IV and class III dendrites. (A) 3rd instar larval expressing GFP-tagged membrane protein (ppk-cd4-tdGFP ) in class IV neurons. (B) The skeleton of a class IV neuron from a third instar larva with an example of a box used to calculate the hitting probability (blue) and a circle used to calculate the correlation dimension dc (magenta). (C) The skeleton of a class III neuron from a third instar larva with an example of a set of boxes used to compute the box dimension db and lacunarity function Λ. See Definition of Morphometrics and Appendix for details (D) Example hitting probability H versus box size B. (E) The correlation function κ versus the radius R. Fits used to determine the correlation dimension dc plotted in solid lines. See Appendix for details of fits. The curves for dc = 1 and dc = 2 are plotted in blue for reference. (F) The number of boxes needed to cover a neuron versus the box size B. Fits used to determine −db are plotted in solid lines. See Appendix for details of fits. The curves for db = 1 and db = 2 are plotted in blue for reference. (G) Lacunarity function Λ versus box size B for a class IV neuron (black) and a class III neuron (red). 1.2 Class IV dendrites are dense and uniform To understand the morphological basis underlying BH we quantified the density and uniformity of the class IV dendrites during the third instar stage. The fractal dimension df is a commonly used measure of how a branched structure fills space. A solid square, for example, has a df = 2, while a straight line has a df = 1 (See Appendix for mathematical definitions of df). We found that the fractal dimension of class IV dendrites was df = 1.80± 0.04 (mean± SD, n = 14), indicating that the dendrites are dense and space filling. To establish a small mesh, a dendrite needs to not only be space filling (i.e. df ∼ 2), but uniformly so. To measure the uniformity of the dendritic arbor, we used the lacunarity function Λ(B). This measures the density variation as a of function length scale B. To compare the lacunarity of different cells we calculated the length BΛ (Λ(B = BΛ) = 0.25 see Definition of Morphometrics). The smaller BΛ, the more uniform the dendritic density. We found that BΛ = 32.6 ± 16.8 µm (mean ± SD, n = 14) for third instar larva. In other words, on spatial scales larger than 33 µm the density of the neuron is uniform, whereas below 33 µm it is variable. While the BΛ is larger than the mesh size BH, it is much smaller than the dendrite size, indicating that the coverage is uniform and arbors can be considered homogeneous. In summary (Tab. 1), mature class IV neurons have dense (large df) and uniform (small BΛ) dendritic arbors, which is consistent with a small mesh size (small BH). 3 BBBCDEFG010020000.250.50.751HittingProbability(H)BoxSize(B)[µm]100 102 104 NumberofBoxesBoxSize(B)[µm]010020000.20.4Lacunarity(Λ)BoxSize(B)[µm]500 µm30 µm30 µmA10.110100100010.1101001000100 103 10CorrelationFunction(κ)Radius(R)[µm]61Class IVClass IIIClass IVClass IIIClass IVClass IIIClass IVClass III 1.3 Class IV dendrites have a tighter mesh, are denser and more uniform than class III dendrites To assess the ability of these measures to quantify dendritic morphology, we also imaged class III cells (Fig. 1 C). Class III cells are gentle touch sensors and use a different set of mechanotransduction channels [25]. The class III dendrites (Fig. 1 C) are substantially less branched and have smaller branches, on average, than the class IV dendrites (Tab 1). We find that H(B) is much smaller for class III neurons than for class IV neurons (Fig. 1 D) for most box sizes. Consequently, we find that the mesh size BH is much larger in class III neurons (BH ∼ 24.7 µm) than in class IV neurons (BH ∼ 8.4 µm) (Tab. 1). In other words, class III dendrites have larger gaps in coverage than class IV dendrites. Table 1: Properties of class IV and class III neurons. All numbers are mean ± SD. Mean branch length, (cid:104)L(cid:105) Mesh size, BH Fractal dimension (correlation method), dc Fractal dimension (box method), db Lacunarity length, BΛ class IV (n = 14) 12.54 ± 0.04 µm 8.37 ± 1.80 µm 1.80 ± 0.04 1.79 ± 0.04 32.6 ± 16.8 µm class III (n = 8) 7.49 ± 0.8 µm 22.7 ± 4.8 µm 1.42 ± 0.03 1.43 ± 0.04 130.5 ± 31.2 µm We find that class III dendrites have a fractal dimension of df ∼ 1.42 (Tab 1), which is substantially smaller than the class IV neuron. In other words, class III dendrites are sparser, at all scales, than class IV dendrites (Fig. 1 E and F). We find that the lacunarity Λ decays much slower (with B) for class III dendrites compared to class IV dendrites (Fig. 1 G). Consequently BΛ is much larger (Tab. 1) for class III dendrites than for class IV neurons (Tab. 1) showing that they are less uniform than class IV neurons. These results (Tab. 1), show that class III neurons are sparser (small df) and less uniform (large BΛ) than class IV dendrites, which is consistent with having larger mesh size (large BH). These differences likely reflect the different mechanoreceptor functions of class III and class IV neurons (see Discussion). 1.4 Class IV dendrites maintain a tight mesh, high density and uniformity throughout larval stages To determine how the morphology of class IV dendrites changes with time, we imaged and skeletonized the class IV dendrites, as before, from early first instar (30 hrs AEL) to wandering third instar (126 hrs AEL) (Fig. 2 A). We used the larval body segment length as a proxy for time since each data point comes from a unique larva, and S increases linearly with time [26] (SI). The mesh size BH increased modestly from 4.4 ± 0.3 µm (1st instar) to 8.5 ± 1.4 µm (third instar) (Fig. 2 C). This two-fold increase is less than the six-fold growth in larval body segment size and indicates that the mesh size remains small during development. Interestingly, the ratio between BH and the mean branch length (cid:104)L(cid:105) is nearly constant during development showing that the shape of the arbor has remarkable conservation during development The morphologies of the class IV neurons remained dense and uniform during the larval stages of development. We found that the fractal dimension (Fig. 2 D) in the just hatched larvae (∼ 30hrs AEL) was df ∼ 1.7 and increased to 1.75 within 24 hours, eventually rising to about 1.8 during the next four days. The lacunarity slightly increased with time (Fig. 2 E). By rescaling B by the radius of gyration Rg (i.e. overall neuronal size see Eq. 1), we found that Λ follows the same curve at all developmental times, i.e., collapses when scaled by Rg (Fig. 2 D inset). The nearly constant fractal dimension and the collapse of the lacunarity curves indicate that the morphological pattern of the class IV neurons is mostly established during embryogenesis. 2 Dynamic Model of Class IV Development The maintenance of a small mesh size throughout larval stages raises the question: how can the class IV dendrites achieve this, despite the six-fold increase in the larval segment size? To answer this question, we developed a mathematical model of class IV dendrite morphogenesis during the larval stages. Our model consists of three rules that determine 1) branch creation, 2) the direction and speed that this new branch grows, and 3) how branches avoid crossing the other branches (self-avoidance). 4 Figure 2: Morphometrics of class IV neurons during larval growth. (A)Examples of class IV neurons during larval stages. All scale bars 30 µm, all time stamps hours After Egg Lay. (B) The hitting probability H is plotted versus the box size B for the 5 neurons shown in Fig. 2 A. In the inset, we plot H versus B/(cid:104)L(cid:105). We define the mesh size BH such that H(B = BH) = 0.5. BH ∼ 0.72(cid:104)L(cid:105) throughout the larval stages. (C) BH is plotted versus larva body segment length S (red 28 cells). For −1 and α = 102 µm (blue). In the inset, we have plotted the mean simulated neurons, ω = 0.2 min branch length of the class IV dendrites versus the larva body segment length S. (D) The fractal dimension df is plotted against S. We have binned the data by body segment length S with bin widths of 77 µm. Simulation parameters are same as before. (E) The lacunarity Λ is plotted against box size B for the five neurons in Fig. 2. In the inset, we plot Λ versus B/Rg. We define the lacunarity length BΛ as the box size at which Λ(B = BΛ) = 0.25. 2.1 Branch Creation We found that the number of branch points increases with time during larval stages (Fig. 3A), and this in- crease was well described by a linear function with a net branch creation rate ωnet ∼ 0.1 branch points min (Tab. 2). This observation led us to model branch creation as a time invariant process with a branching frequency ω. Since ωnet can include the removal of branches (see below), it is a lower bound on the branching creation frequency ω. −1 We also measured the branch length distribution, and found that it was well described be an expo- nential distribution (Fig. 3 B). Furthermore, we found that that the branch length was independent of the branch depth, defined as the number of branch points between a branch and the soma along the shortest path (Fig. 3 C). Motivated by these observations, we modeled branch creation as a random process that was uniform along the neuron and constant in time. 2.2 Tip Elongation Since the growth of class IV dendrites occurs at the branch tips [27, 6], we modeled neuron growth as branch tip elongation. We measured the overall size of class IV dendrites using the radius of gyration Rg (Eq. 1). Rg measures spread of neuron mass from its center. We found that Rg increases linearly during development (Fig. 3 D). Therefore, we assume that the tip elongation rate v is constant in time and v ∝ Vg, where Vg is the growth speed of the class IV dendrite. The assumption of a constant growth speed assumes that the simulation time scale is much larger than the fluctuation times scale. We estimated Vg −1 (Tab 2). from the change in dendrite size during development (Fig. 3 D), and found Vg ∼ 0.04 µm min To determine the direction of branch growth, we measured how much the path of a branch changes as a function of path length by using the persistence length β; the distance over which the orientation 5 020040000.20.4Lacunarity(Λ)BoxSize(B)[µm]012B/Rg00.20.4ΛBΛ/Rg30 hr56 hr80 hr96 hr126 hr02004006001.61.71.8FractalDimension(df)SegmentLength(S)[µm]EC05010000.250.50.751HittingProbability(H)BoxSize(B)[µm]0123B/(cid:31)L(cid:30)00.250.50.751HBH/(cid:31)L(cid:30)30 hr56 hr80 hr90 hr126 hrBA0200400600051015SegmentLength(S)[µm]BH[µm]0200400600051015(cid:31)L(cid:30) [µm] S [µm]30 hrs56 hrs80 hrs96 hrs126 hrsDdbdcSimulation30 µm Figure 3: Branching Rules (A) Number of branches versus larval body segment length S for class IV dendrites at different larval stages. The slope of the fit is 1.64 branches µm−1 for the line passing through the S axis at 50 µm, the segment length at the onset of dendritogenesis. (B) The probability density P (L) that a branch of a dendrite (at a particular time) has length L. The probability density is rescaled by multiplying by (cid:104)L(cid:105) and the branch length is rescaled by dividing by (cid:104)L(cid:105) (of the particular dendrite). The superposition of the data indicates that the branch lengths are well described be an exponential distribution (red line) at all larval stages. (C) The branch length autocorrelation Cl (Eq. 3) vs depth difference ∆d averaged over all branches. The near-zero values for ∆d ≥ 1 imply that branch length is independent of depth. (D) The radius of gyration Rg (Eq. 1) of 28 class IV neurons versus larval body segment length S. The slope of the fit is 0.39, where the S−intercept is set such that Rg = 0 at the onset of class IV dendrite morphogenesis. (E) The tangent vector autocorrelation function Ct (Eq. 2), averaged over all branches, versus the path length lag ∆s (Eq. 2). The red line is an exponential fit to the data Ct(∆s) = e∆s/β, where β = 44.8 ± 1.5 µm is the persistence length. of the direction of growth of a branch changes (see Appendix for mathematical details). We find that β = 45 ± 2 µm (Tab. 2, Fig. 3 E), which is much greater than the mean branch length ((cid:104)L(cid:105) ∼ 12.5 Tab. 1), indicating that branches tend to be straight. 2.3 Self-Avoidance Previous work has shown that the branches of the class IV neuron do not cross each other. Furthermore, it is known that this self-avoidance is contact mediated and it has been proposed that branches retract after contact [27, 28, 29, 30]. Therefore, we modeled self-avoidance by having growing branches retract at a constant speed v (same as elongation rate), if they contact an existing branch. For simplicity, the retraction length was assumed to be exponentially distributed with a mean retraction length α. Table 2: Parameters in the model. All errors are SD (n=28 neurons). Parameter Branching frequency Tip elongation rate, v Persistence length, β Retraction length, α 0.12 ± 0.03 min 0.04 ± 0.02 µm · min −1 (ωnet) −1 Measured Value 44.8 ± 1.5 µm not measured Simulation Value −1(ω) 0.01 − 2 min −1 0.08 µm · min 70 µm 0 − 1000 µm 2.4 Model Implementation Using these rules, we arrived at a four parameter model; ω the branch creation rate, v the growth speed, β persistence length and α to retraction scale. We implemented our model on a hexagonal lattice, with lattice spacing  = 0.4 µm. We chose to use a lattice model to facilitate contact detection as the lattice spacing was set to be equal to the thickness of the branches. Since we implemented the rules on a hexagonal lattice, the possible branching angles are limited. Therefore we ignored the role of branching angles on morphology. The lattice spacing and the growth speed act as scale factors, i.e. setting the overall size of the dendrite but not changing the shape. Therefore, we chose v such that the size of a simulated dendrite matched that of a real one, which left us with three free parameters. However, we found that as long as β is large, i.e., as long as the branches are straight, β has little effect on the morphology (SI). Therefore, we focused on how branching frequency ω and the contact-based retraction scale α control the morphology of dendrites. 6 03006000150300Segment Length (S) [µm]R g [µm]024600.51Depth (∆d)05101500.51Path Length (∆s) [µm]Ct(∆s)Cl(∆d)0240.010.11L/(cid:31)L(cid:30)P(L)(cid:31)L(cid:30)030060005001000Number of Branch PointsSegment Length (S) [µm]ABCDEDataFitDataFitDataFitDataFit 2.5 Model Results Figure 4: Simulations of dendritic growth (A) Nine representative simulations of dendrties for a range of branching frequencies ω and retraction scales α. For all simulations, the persistence length was −1 (dashed box) we β = 70 µm and simulation time was T = 100 hrs. For α (cid:29) 1 µm and ω ∼ 0.2min have subjective agreement with the morphology of real neurons. (B) The branch length distribution for simulated neurons in the dashed black box in (A) is approximately exponentially distributed as observed. (C) The mesh size BH in a contour plot versus the branching frequency ω and retraction scale α. Note that the retraction scale α is plotted in a log scale. In pink we have highlighted the observed value of BH = 7 µm (see Tab. 1). The white box indicates the values used for the time series −1 and α = 102 µm). (D) The fractal dimension df in a contour plotted in Fig. 2 C and D (ω = 0.2 min plot versus the branching frequency ω and the retraction scale α. In pink we have highlighted the physiological value of df = 1.8 (see Tab. 1) (E) The lacunarity scale BΛ in a contour plot versus the branching frequency ω and the retraction scale α. We highlighted in pink the observed value BΛ = 33 µm (see Tab. 1). We find that this value is not found in the range of values used in our simulations. We generated simulated neurons for a range of branching frequencies ω and retraction scales α (Fig. 4 A) and analyzed their shape. Importantly, we found that the distribution of branch lengths was exponential (Fig. 4 B) and the branch lengths were uncorrelated with branch depth, for a limited range of parameters (dashed box in Fig. 4 A), which agrees with the experimental observations (Fig. 3 B and C). To test whether our model provides a good description of the morphology, we measured the mesh size, fractal dimension and lacunarity of the simulated dendrites. The mesh size BH decreased with increasing ω and decreasing α (Fig. 4 C) and appeared to saturate as ω increases. There was a small region of ω− α space where there was quantitative agreement between the simulated and observed values of BH (Fig. 4 C, pink). We found, that df increased monotonically with ω and saturated for large values of ω. As for BH, we found a narrow band of ω − α space where we have a quantitative agreement between the model and the class IV neuron for df (Fig. 4 D, pink). Crucially, there was a small region of ω − α space where the model recapitulated both BH and df (Fig. 4 C and D, white box). Taking values from this small −1 and α = 102 µm), we recapitulated the third instar and even found agreement region (ω = 0.2 min throughout most of the larval development (Fig. 2 B and 2 C). Thus, the model recapitulates both the mesh size and fractal dimension throughout development. It did not recapitulate the lacunarity, which −1 and α = 102 µm was larger than the observed values. In other for the parameter values ω = 0.2 min words, the model arbors are not as uniform as the observed ones (see Discussion). In summary, this model recapitulates most, but not all aspects of the dendritic morphology of class IV neurons. 7 Branching Frequency (ω) [min-1]10240.010.11P(L)(cid:31)L(cid:30)L/(cid:31)L(cid:30)Retraction Scale (α) [µm]00.51.52110010110210360120180240BΛ [µm] Branching Frequency (ω) [min-1]Retraction Scale (α) [µm]00.51.5211001011021031.71.8Fractal Dimension (dc)Branching Frequency (ω) [min-1]Retraction Scale (α) [µm]00.51.52110010110210336912BH [µm]Branching Frequency (ω) [min-1]Retraction Scale (α) [µm]0.020.22101001000CBDEA 3 Discussion Our key experimental finding is that the morphology of class IV neurons, as characterized by the branch length, mesh size, fractal dimension, and lacunarity, remarkably remains constant during development. Indeed, from the early first instar larva (30 hours after egg lay) to the late third instar larva (126 hours after egg lay), as both the segment size and the number of branches increases approximately six-fold, the mean length, the mesh size, and the lacunarity only increase around two-fold and the fractal dimension is almost unchanged. When we normalize the mesh size by the mean branch length, it is virtually unchanged throughout development. Thus, as these cells grow, key aspects of their morphology are invariant, even as the overall size the number of branches increases by nearly an order of magnitude. As a first step toward probing the mechanism underlying this geometric invariance, we developed a simple computational model for branched morphogenesis. The model assumes that the branching rate is constant over development (consistent with the observed linear increase in the number of branches), that the rate was independent of position and that growing tips retract random, exponentially distributed distance after contacting other branches. The model reproduced many of the key features of the growth of class IV cells including branch lengths, mesh sizes, and fractal dimensions. However, it was unable to capture the lacunarity (the model predicted a higher relative density in the center than was observed, see SI). Importantly, the data constrained the values of the branching frequency and mean retraction distance. Thus, the model provides a framework for understanding the changes in the morphology of these cells during development. One of our most striking experimental and theoretical findings was that the branching frequency in class IV dendrites was independent of total dendrite length. Naively, we might have expected that the mean number of branches added per unit time would increase with total dendrite length, as the longer the dendrites, the more positions on which branches could form. However, this would have led to an exponential increase in the number of branches, rather than the observed linear increase. Our modeling shows that even if the retraction length is much larger than the mean branch length, an exponential increase in branch length is still observed, and the distribution of branch lengths deviates from the observed exponential distribution (see SI). The constant branching rate suggests that branching is limited by the production of a nucleating factor that is produced at a constant rate. Furthermore, our finding that branching is uniform in space implies that the putative nucleation factor would be distributed widely and uniformly throughout the cell. We also found that for our model to recapitulate class IV-like morphologies, contact-based retraction needs to lead to complete branch deletion, i.e., the mean retraction distance (α) is much larger than the mean branch length ((cid:104)L(cid:105)). By deleting the branches whose tips collide with other branches, gaps of a size similar to the mean branch length are created and maintained. Also, dense regions where the gap size is less than or equal to the mean branch length will not increase in density. Such overfilling is seen in the simulations for small retraction lengths (Fig. 4 A). Thus, our model constrains both the branching frequency and the retraction distance. Finally, we note that the mesh size of class IV dendrites, 4 − 8 µm, is well suited for detecting highly localized nociceptive stimuli such as punctures by the 10 µm diameter ovipositor barb of parasitic wasps ([8]). This acuity is maintained throughout development. Thus, the small mesh size is consistent with the class IV neuron being a harsh touch sensor. Indeed, the theory of contact dynamics predicts that the indentation h of the surface of an elastic body poked by a probe with a cross-sectional radius R (pushing normal to the surface) is h ∝ F/R, where F is the applied force [31]. Therefore, the smaller R, the larger h (for a fixed force); the more local the stimuli, the more sensitive to localized forces. Thus the small mesh size suggests that the class IV neuron is well adapted to sensing harsh touch throughout larval stages. In contrast, the class III neuron, a soft touch sensor, would need to capture diffuse stimuli, i.e., the mesh size can be large. Thus, the morphologies of both class IV and class III neurons are well-suited for their mechanosensitive functions. 4 Materials and Methods Drosophila Stocks All flies were maintained on standard medium at 23oC. The strain ppk-cd4-tdGFP was a kind gift from Dr. Han Chun (Cornell University). 8 Imaging and Skeletonization The larvae were mounted in 50% glycerol in PBS between a glass slide and a cover slip. The sam- ple was imaged using a confocal laser scanning microscope (Zeiss, LSM780) with 63x objective. The 600x600 µm images were stitched together offline using Fiji and the stitched images were processed using the NeuronStudio [23] to obtain the skeleton a one pixel wide tracings of the dendritic arbors. A Mathematical Definitions of Morphometrics Radius of gyration of the Neuron The radius of gyration is defined as Rg = (cid:118)(cid:117)(cid:117)(cid:116) 1 M M(cid:88) j=1 (rj − rm)2, (1) where M is the total number of occupied pixels, rj is the position of the jth occupied pixel and rm is the mean position of all occupied pixels. Rg measures the standard deviation of the dendrite pixels, i.e., the spread of the imaged neuron from its center. Path Correlation and Persistence Length The deviation of a branch path from a straight line can be quantified using the tangent vector autocor- relation function Ct(∆s) ∼ (cid:104)t(s) · t(s + ∆s)(cid:105)s, (2) where t(s) is the tangent vector as a function of the path length s. Ct measures the angular change of the t as a function of path length, i.e., how bent the branch is. If Ct = 1, the path is straight and if Ct = 0, there is a 90o turn in the path. Branch Length Correlation Function The branch length autocorrelation function is Cl(∆d) = (cid:104)L(d)L(d + ∆d)(cid:105)d − (cid:104)L(d)(cid:105)2 d (cid:104)L2(d)(cid:105)d − (cid:104)L(d)(cid:105)2 d , (3) where L(d) is the branch length at depth d and ∆d is the depth difference. Depth is defined as the number of branch points between the branch and the soma, along the shortest path from the branch to the soma. (cid:104). . .(cid:105)d represents the average over d. Hitting Probability Consider a box with side length R centered anywhere in the receptor field of the neuron (Fig. 1 C). We then ask: 'what is the probability PH (b, R) of having b pixels in a box of size R?'. Using this probability, we can determine the probability that a box of size R contains at least n pixels (cid:90) M Hn(R) = PH (b, R)db, n (4) where M is the total number of neuron pixels. We define the mesh size BH such that H1(R = BH) = 0.5, i.e., the mesh size is the box width such that there is a 50% chance that the box contains at least one pixel from the skeleton. 9 Fractal Dimension In this paper, the fractal dimension is measured using two different methods: the correlation dimension (Fig. 1 E) and box counting (Fig. 1 F) method. In the box counting method, we determine the number of boxes N of side length R that are needed to cover the neuron (Fig. 1 B). The number of boxes needed to cover a line of length l is N = l R ; therefore N ∝ R−1. The number of boxes needed to cover a square ; therefore N ∝ R−2. In general, N ∝ R−db, where db is the box counting of side length l is N =(cid:0) l measure of the fractal dimension. (cid:1)2 R In the correlation method [32], we determine how many pixels are contained within a circle of radius R. Let each point xi on the neuron be the center of a circle of radius R (Fig. 1 B). Then N (xi, R) is the number of skeleton pixels in the circle (green pixels in Fig. 1 C). Averaging over all possible centers (i.e. skeleton pixels) xi, gives κ(R) = (cid:104)N (xi, R)(cid:105)xi. (5) In general, κ(R) ∝ Rdc where dc is the correlation measure of the fractal dimension. The relation f (R) ∝ Rd is called a scaling law and is only valid in a finite range of R (e.g. for small R we approach the scale of one pixel, and for large R we approach the total dimension of the neuron). For the neurons the minimum scale is half the mean branch length and the maximum scale is the radius of gyration. Lacunarity Consider the set of boxes of linear dimension R used in the box counting method (see figure 1 B). Instead of asking how many boxes are needed to cover the shape, we ask 'what is the probability PB(b, r) of having b pixels in a box of size R?'. PB(b, r) differs from PH (b, R) since it only considers boxes that have at least one pixel (b ≥ 1). The moments of PB are defined as (cid:90) M µn(R) = 1 bnPB(b, R)db, where M is the total number of skeleton pixels. This then allows us to look at the coefficient of variation of PB(b, R) CV (R) = σ2 µ2 1 = µ2 − µ2 1 µ2 1 , (6) where σ2 is the variance of PB(b, R). CV (R) is also called the lacunarity function. The more uniform a shape is, the smaller CV and the more variable the shape, the larger the CV . For example, a uniform shape would have a CV ∼ 0. How large CV needs to be for a shape/neuron to be consider variable is somewhat arbitrary. The more important point is that the larger CV , the larger the variation in the neuron. It also allows us to see how these variations change with length scale. Thus, the lacunarity function measures the variations and assigns them a typical length scale. Acknowledgments This work was partially supported by the NIH Pioneer Award (Award Number), National Natural Science Foundation of China (NSFC Grant 31671389, to X.L.) and Max-Planck Partner Program (to X.L.). SG was supported by an EMBO Long-Term Fellowship, and OT is supported by the Fonds de Reshershe du Qubec - Nature et technologies. Thanks to Dr. Han Chun (Cornell University) for the fly strains. References [1] Greg Stuart, Nelson Spruston, and Michael Hausser, editors. Dendrites. Oxford University Press, 2nd edition, 2007. [2] D Bentley and A Toroian-Raymond. Embryonic and postembryonic morphogenesis of a grasshopper interneuron. J Comp Neurol, 201(4):507–18, Oct 1981. 10 [3] J W Truman and S E Reiss. Hormonal regulation of the shape of identified motoneurons in the moth manduca sexta. J Neurosci, 8(3):765–75, Mar 1988. [4] Yan Li, Diana Brewer, Robert E Burke, and Giorgio A Ascoli. Developmental changes in spinal motoneuron dendrites in neonatal mice. J Comp Neurol, 483(3):304–17, Mar 2005. [5] P F Hitchcock. Constant dendritic coverage by ganglion cells with growth of the goldfish's retina. Vision Res, 27(1):17–22, 1987. [6] Jay Z Parrish, Peizhang Xu, Charles C Kim, Lily Yeh Jan, and Yuh Nung Jan. The microrna bantam functions in epithelial cells to regulate scaling growth of dendrite arbors in drosophila sensory neurons. Neuron, 63(6):788–802, Sep 2009. [7] Rolf Bodmer and Yuh Nung Jan. Morphological differentiation of the embryonic peripheral neurons in drosophila. Roux's archives of developmental biology, 196(2):69–77, 1987. [8] Richard Y Hwang, Lixian Zhong, Yifan Xu, Trevor Johnson, Feng Zhang, Karl Deisseroth, and W Daniel Tracey. Nociceptive neurons protect drosophila larvae from parasitoid wasps. Curr Biol, 17(24):2105–16, Dec 2007. [9] H Wassle and B B Boycott. Functional architecture of the mammalian retina. Physiol Rev, 71(2):447–80, Apr 1991. [10] M Hausser, N Spruston, and G J Stuart. Diversity and dynamics of dendritic signaling. Science, 290(5492):739–44, Oct 2000. [11] D Attwell and S B Laughlin. An energy budget for signaling in the grey matter of the brain. J Cereb Blood Flow Metab, 21(10):1133–45, Oct 2001. [12] Gordon M G Shepherd, Armen Stepanyants, Ingrid Bureau, Dmitri Chklovskii, and Karel Svoboda. Geometric and functional organization of cortical circuits. Nat Neurosci, 8(6):782–90, Jun 2005. [13] Nelson Spruston. Pyramidal neurons: dendritic structure and synaptic integration. Nat Rev Neu- rosci, 9(3):206–21, Mar 2008. [14] Quan Wen and Dmitri B Chklovskii. A cost-benefit analysis of neuronal morphology. J Neurophysiol, 99(5):2320–8, May 2008. [15] Yuh-Nung Jan and Lily Yeh Jan. Branching out: mechanisms of dendritic arborization. Nat Rev Neurosci, 11(5):316–28, May 2010. [16] S Lawrence Zipursky and Wesley B Grueber. The molecular basis of self-avoidance. Annu Rev Neurosci, 36:547–68, Jul 2013. [17] Xintong Dong, Kang Shen, and Hannes E Bulow. Intrinsic and extrinsic mechanisms of dendritic morphogenesis. Annu Rev Physiol, 77:271–300, Feb 2015. [18] Wesley B Grueber, Lily Y Jan, and Yuh Nung Jan. Tiling of the drosophila epidermis by multiden- dritic sensory neurons. Development, 129(12):2867–78, Jun 2002. [19] Jessica L Robertson, Asako Tsubouchi, and W Daniel Tracey. Larval defense against attack from parasitoid wasps requires nociceptive neurons. PLoS One, 8(10):e78704, 2013. [20] Benoit Mandelbrot. How long is the coast of britain? statistical self-similarity and fractional dimension. Science, 156(3775):636–638, 1967. [21] Benoit B. Mandelbrot. The fractal geometry of nature. W.H. Freeman, 1 edition, August 1982. [22] C. Allain and M. Cloitre. Characterizing the lacunarity of random and deterministic fractal sets. Phys. Rev. A, 44:3552–3558, Sep 1991. [23] S L Wearne, A Rodriguez, D B Ehlenberger, A B Rocher, S C Henderson, and P R Hof. New techniques for imaging, digitization and analysis of three-dimensional neural morphology on multiple scales. Neuroscience, 136(3):661–80, 2005. 11 [24] Masao Doi and Sam F Edwards. The theory of polymer dynamics, volume 73. oxford university press, 1988. [25] Zhiqiang Yan, Wei Zhang, Ye He, David Gorczyca, Yang Xiang, Li E Cheng, Shan Meltzer, Lily Yeh Jan, and Yuh Nung Jan. Drosophila nompc is a mechanotransduction channel subunit for gentle- touch sensation. Nature, 493(7431):221–5, Jan 2013. [26] Michael Ashburner, Kent G Hawley Golic, et al. Drosophila: a laboratory handbook. Technical report, 2005. [27] Benjamin J Matthews, Michelle E Kim, John J Flanagan, Daisuke Hattori, James C Clemens, S Lawrence Zipursky, and Wesley B Grueber. Dendrite self-avoidance is controlled by dscam. Cell, 129(3):593–604, May 2007. [28] Peter Soba, Sijun Zhu, Kazuo Emoto, Susan Younger, Shun-Jen Yang, Hung-Hsiang Yu, Tzumin Lee, Lily Yeh Jan, and Yuh-Nung Jan. Drosophila sensory neurons require dscam for dendritic self-avoidance and proper dendritic field organization. Neuron, 54(3):403–16, May 2007. [29] Katie M Hutchinson, Fernando Vonhoff, and Carsten Duch. Dscam1 is required for normal den- drite growth and branching but not for dendritic spacing in drosophila motoneurons. J Neurosci, 34(5):1924–31, Jan 2014. [30] Benjamin J Matthews and Wesley B Grueber. Dscam1-mediated self-avoidance counters netrin- dependent targeting of dendrites in drosophila. Curr Biol, 21(17):1480–7, Sep 2011. [31] Ian N Sneddon. The relation between load and penetration in the axisymmetric boussinesq problem for a punch of arbitrary profile. International Journal of Engineering Science, 3(1):47–57, 1965. [32] Peter Grassberger and Itamar Procaccia. Measuring the strangeness of strange attractors. Physica D: Nonlinear Phenomena, 9(1–2):189 – 208, 1983. 12
1012.1611
2
1012
2011-03-05T17:28:20
Broken chaotic clocks of brain neurons and depression
[ "q-bio.NC", "cond-mat.dis-nn", "nlin.CD" ]
Irregular spiking time-series obtained in vitro and in vivo from singular brain neurons of different types of rats are analyzed by mapping to telegraph signals. Since the neural information is coded in the length of the interspike intervals and their positions on the time axis, this mapping is the most direct way to map a spike train into a signal which allows a proper application of the Fourier transform methods. This analysis shows that healthy neurons firing has periodic and chaotic deterministic clocks while for the rats representing genetic animal model of human depression these neuron clocks might be broken, that results in decoherence between the depressive neurons firing. Since depression is usually accompanied by a narrowing of consciousness this specific decoherence can be considered as a cause of the phenomenon of the consciousness narrowing as well. This suggestion is also supported by observation of the large-scale chaotic coherence of the posterior piriform and entorhinal cortices' electrical activity at transition from anesthesia to the waking state with full consciousness.
q-bio.NC
q-bio
Broken chaotic clocks of brain neurons and depression ICAR, P.O.B. 31155, Jerusalem 91000, Israel A. Bershadskii Irregular spiking time-series obtained in vitro and in vivo from singular brain neurons of different types of rats are analyzed by mapping to telegraph signals. Since the neural information is coded in the length of the interspike intervals and their positions on the time axis, this mapping is the most direct way to map a spike train into a signal which allows a proper application of the Fourier transform methods. This analysis shows that healthy neurons firing has periodic and chaotic de- terministic clocks while for the rats representing genetic animal model of human depression these neuron clocks might be broken, that results in decoherence between the depressive neurons firing. Since depression is usually accompanied by a narrowing of consciousness this specific decoherence can be considered as a cause of the phenomenon of the consciousness narrowing as well. This sug- gestion is also supported by observation of the large-scale chaotic coherence of the posterior piriform and entorhinal cortices' electrical activity at transition from anesthesia to the waking state with full consciousness. PACS numbers: 87.19.L, 87.19.ll, 87.19.lm I. INTRODUCTION All types of information, which is received by sen- sory system, are encoded by nerve cells into sequences of pulses of similar shape (spikes) before they are trans- mitted to the brain. Brain neurons use such sequences as main instrument for intercells connection. The infor- mation is reflected in the time intervals between succes- sive firings (interspike intervals of the action potential train, see Fig. 1). There need be no loss of information in principle when converting from dynamical amplitude information to spike trains [1] and the irregular spike se- quences are the foundation of neural information process- ing. Although understanding of the origin of interspike intervals irregularity has important implications for elu- cidating the temporal components of the neuronal code and for treatment of such mental disorders as depression and schizophrenia, the problem is still very far from its solution. The mighty Fourier transform method, for instance, is practically non-applicable to the spike time trains. The spikes are almost identical to each other and the neural information is coded in the length of the interspike in- tervals and the interspike intervals positions on the time axis, therefore it is the most direct way to map the spike train into a telegraph time signal, which has values -1 from one side of a spike and values +1 from another side of the spike with a chosen time-scale resolution. An ex- ample of such mapping is given in figure 1. While the information coding is here the same as for the corre- sponding spike train, the Fourier transform method is quite applicable to analysis of the telegraph time-series. On the other hand, recent dynamical models of neu- ron activity revealed new and complex role of regimes of a (deterministic) chaotic irregularity in the neuron spike trains (see, for instance, [2]-[5]). Therefore, we have to use all available mathematical tools in order to study the experimental data on the deterministic chaos properties (and, especially, in order to separate between determin- h p a r g e e t l s e k p s i 1 0 -1 1 0 0 a b 5 10 15 20 25 t [s] FIG. 1: Mapping of a spike train (figure 1b) into a telegraph signal (figure 1a). istic chaos and stochasticity in the experimental signals). In present paper we have analyzed three types of exper- imentally obtained spike trains: a) obtained in vitro from a spontaneous activity in CA3 hippocampal slice culture of a healthy Wistar/ST rat (the raw data and the de- tail description of the experiment can be found online at http://hippocampus.jp/data and in Refs. [6],[7]), b) ob- tained in an electrophysiological in vivo experiment from neurons belonging to red nucleus of a healthy (Sprague- Dawley) rat's brain (see Ref. [8] for more details of the experiment and a preliminary discussion of the data), and c) obtained in an electrophysiological in vivo experiment from neurons belonging to red nucleus of a genetically depressed (Flinders Sensitive Rat Line) rat's brain (see Ref. [8] for more details and a preliminary discussion of the data). In the in vitro experiment a) a functional imaging technique with multicell loading of the calcium fluo- rophore was used in order to obtain the spike trains of ) f ( E 25 20 15 10 5 0 0 0.5 1.5 2 1 f [s-1] 100 10 1 0.1 ) f ( E 2 -1.7 0 2 4 6 8 10 12 14 16 τ [s-1] 0.01 0.01 0.1 1 10 f [s-1] ) τ ( C 1 0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 FIG. 2: Autocorrelation function for the telegraph signal cor- responding to the cell-21 (800 spikes). Insert in the Fig. 2 shows corresponding spectrum. FIG. 3: Spectrum of the telegraph signal corresponding to the cell-21 in log-log scales. The dashed straight line indicates a power law: E(f ) ∼ f −1.7. spontaneously active singular neurons in the absence of external In the in vivo experiments b) and c) the rats were anesthetized and the extracel- lular recordings were processed from the singular cells [8]. input [6],[7]. Motivation to study the hippocampus and red nucleus areas of brain in relation to the psychomotor aspects of depression is based on the recently discovered evidences of their deep involvement in this mental disorder. The hippocampus is a significant part of a brain system re- sponsible for behavioral inhibition and attention, spa- tial memory, and navigation. It is also well known that spatial memory and navigation of the rats is closely re- lated to the rhythms of their moving activity. On the other hand, the hippocampus of a human who has suf- fered long-term clinical depression can be as much as 20% smaller than the hippocampus of someone who has never been depressed [9]. Inputs to the Red Nucleus arise from motor areas of the brain and in particular the deep cere- bellar nuclei (via superior cerebellar peduncle; crossed projection) and the motor cortex (corticorubral; ipsilat- eral projection). On the other hand it is known that humans with deep depression have intrinsic locomotors problems. Therefore, investigation of Red Nucleus for genetically defined rat model of depression (these rats partially resemble depressed humans because they ex- hibit reduced appetite and psychomotor function) can be useful for understanding the mental disorder origin. II. NEURON CLOCK In the in vitro experiment with spontaneous activ- ity of the hippocampal pyramidal cells different levels of activity were observed for different neurons [6],[7]. We take for our analysis the two most active neurons (http://hippocampus.jp/data - Data-006, cell-21, with 800 spikes in the time-series; and cell-25, with 692 spikes). 100 10 1 0.1 ) f ( E -1.7 0.01 0.01 0.1 1 10 f [s-1] FIG. 4: As in Fig. 3 but for cell-25. The spike trains were mapped to telegraph signals as it is described above (see also [10]). Figure 2 shows autocor- relation function for the telegraph signal corresponding to the cell-21 (800 spikes). Insert in the Fig. 2 shows corresponding spectrum. Both the correlation function and the spectrum provide clear indication of a strong periodic component in the signal (the oscillations in the correlation function and the peak in the spectrum). The periodic component can be seen at frequency f0 ≃ 0.3Hz. Figure 3 shows the spectrum in log-log scales. One can see that at high frequencies the spectrum exhibits a scal- ing behavior (power law: E(f ) ∼ f −1.7, as indicated by the dashed straight line). The real power law can be more pronounced but under the experimental conditions indi- vidual spikes emitted at firing rates higher than 5Hz were experimentally inseparable [6],[7]. Figure 4 shows spec- trum of the telegraph signal corresponding to the spike train obtained for the cell-25 (692 spikes). The spectrum is rather similar to the spectrum shown in Fig. 3 (for cell-21). The more broad peak in Fig. 4 can be related 3 20 15 10 5 0 -5 X -10 -15 0 threshold 2000 4000 t 6000 8000 30 25 20 Z 15 10 5 0 0 50 100 150 t 200 250 300 FIG. 5: X-component fluctuations of a chaotic solution of the Rossler system Eq. (1) (a = 0.15, b = 0.20, c = 10.0). FIG. 6: Z-component fluctuations of a chaotic solution of the Rossler system. to the poorer statistics for the cell-25 in comparison with cell-21. The spectra were calculated using the maximum entropy method (because it provides an optimal spectral resolution even for small data sets [11]). III. THRESHOLD EFFECT Now let us speculate about physics which could result in the spectra observed in Figs. 3 and 4. And let us recall some basic electrochemical properties of neuron. Nerve cells are surrounded by a membrane that allows some ions to pass through while it blocks the passage of other ions. When a neuron is not sending a signal it is said to be "at rest". At rest there are relatively more sodium ions onside the neuron and more potassium ions inside that neuron. The resting value of the membrane electro- chemical potential P (the voltage difference across the neural membrane) of a neuron is about -70mV. If some event (a stimulus) causes the resting potential to move toward 0mV and the depolarization reaches about -55mV (a "normal" threshold) a neuron will fire an action poten- tial. The action potential is an explosive release of charge between neuron and its surroundings that is created by a depolarizing current. If the neuron does not reach this critical threshold level, then no action potential will fire. Also, when the threshold level is reached, an action po- tential of a fixed size will always fire (for any given neu- ron the size of the action potential is always the same). Depending on different types of voltage-dependent ion channels, different types of action potentials are gener- ated in different cells types and the qualitative estimates of the potentials and time periods can be varied. Recent reconstructions of a driver of the membrane potential us- ing the neuron spike trains indicate the Rossler oscillator as the most probable (and simple) candidate (see, for in- stance, Refs. [10],[12]-[17]). Figure 5 shows as example the x-component fluctuations of a chaotic solution of the Rossler system [18] 100 10 1 0.1 ) f ( E -1.7 0.01 0.001 0.01 0.1 1 f FIG. 7: Spectrum of the telegraph signal corresponding to the spike train generated by the x-component fluctuations overcoming the threshold x = 7. The dashed straight line indicates a power law: E(f ) ∼ f −1.7. dx dt = −y − z; dy dt = x + ay; dz dt = b + xz − cz (1) where a, b and c are parameters). At certain values of the parameters a,b and c the z-component of the Rossler system is a spiky time series [19],[20]: Fig. 6. It can be shown that the Rossler system and the well known Hindmarsh-Rose model [21] of neurons are subsystems of the same differential model with a spiky component [20]. Previously the 'spiky' component of such models was in- terpreted and studied as a simulation of a neuronal out- put. For the spontaneous neuron firing (without external stimulus), however, we suggest to reverse the approach and consider the spiky variable as the main component of the electrical input (which naturally should have a 'spiky' character, see above) to the neuron under consideration. For each neuron the height of the spikes, which the neu- ron generates, is about the same. However, the heights E(f) ~ exp-f/fe 4 E(f) ~ exp-f/fe Te=1/fe=0.9 s 3 2 1 0 ) f ( E n l 0 0.1 0.2 0.3 0.4 0.5 f 0 1 3 4 2 f [s-1] ) f ( E n l 10 5 0 -5 -10 -15 -20 FIG. 8: Spectrum of the x-component fluctuations shown in Fig. 5. We used the semi-log axes in order to indicate expo- nential decay of the spectrum. FIG. 9: Spectrum of the telegraph signal corresponding to a healthy red nucleus cell. The data is shown in the semi-log scales in order to indicate the exponential decay Eq. 2 (the straight line). of the spikes generated by different neurons are differ- ent. Also the signals coming from different neurons to the neuron under consideration have to go through the electrochemical passes with different properties. There- fore, the spiky z-time-series (Fig. 6) can naturally rep- resent a multineuron signal, which can be considered as a spontaneous input for the neuron under consideration. If we use the usual interpretation of the x-component as a driver of the membrane potential P (x) and the y- component as that taking into account the transport of ions across the membrane through the ion channels [21], then the position of the input (the component z) in the first equation of the system Eq. (1) has a good phys- ical background (cf. Ref. [21]). Then, the quadratic nonlinearity in the third equation of the system Eq. (1) can be interpreted as a simple (in the Taylor expansion terms) feedback of the neuron to the main component of the neuronal input. This model with the strong non- linear feedback can be relevant to the most active neu- rons of a spontaneously active brain (see below results of an in vitro experiment with a spontaneous brain activ- ity). The details of the function P (x) is not significant for the threshold firing process, what really matters is that the membrane potential function P (x) reaches its firing value when (and only when) its argument x crosses certain threshold from below. In this simple model the driving variable x may overcome its threshold value (Fig. 5) due to the deterministic (chaotic) spontaneous stimu- lus. Let us consider an output spike signal resulting from overcoming a threshold value x = 7, for instance. Fig. 7 shows spectrum of the telegraph signal corresponding to the spike signal. One can compare Fig. 7 with Figs. 3 and 4 to see very good reproduction of the main spectral properties. In order to understand what is going on here we show in figure 8 spectrum of the x-component itself. The semi-log scales are used in these figures in order to indicate exponential decay in the spectra (in the semi-log scales this decay corresponds to a straight line): E(f ) ∼ e−f /fe (2) While the peak in the spectrum corresponds to the funda- mental frequency, f0, of the Rossler chaotic attractor, the rate of the exponentional decay (the slope of the straight line in Fig. 8 provides us with and additional charac- teristic frequency fe. Thus Rossler chaotic attractor has two clocks: periodic with frequency f0 and chaotic with frequency fe. If one compares Fig. 8 and Fig. 7 one can see that the periodic clock survived the threshold cross- ing (with a period doubling, see Appendix). The chaotic clock, however, did not survive the threshold crossing at spontaneous activity: the exponential decay in Fig. 8 has been transformed into a scaling (power law) decay in Fig. 7, which has no characteristic frequency (scale invari- ance). The scaling exponent value '-1.7' is not sensitive to a reasonable variation of the threshold value (∼ 20%) and even to Gaussian fluctuations of the threshold value. Therefore, it is not just a coincidence that the scaling law in the Rossler case agrees with results of the in vitro ex- periment (cf. also [22],[23],[24] and Fig. 16a). It should be noted that for a wide class of deterministic systems a broad-band spectrum with exponential decay is a generic feature of their chaotic solutions Refs. [11],[25]-[27]. It is shown in Ref. [25] that the characteristic frequency fe = X λ+ i i (3) where λ+ system. i are positive Lyapunov exponents of the chaotic ) f ( E n l 4 3 2 1 0 -1 -2 0 E(f) ~ exp-f/fe Te=1/fe=0.9s 1 2 f [s-1] 3 4 FIG. 10: As in Fig. 9 but for another healthy cell. IV. CHAOS VS. STOCHASTICITY IN NEURON FIRING Both stochastic and deterministic processes can result in the broad-band part of the spectrum, but the decay in the spectral power is different for the two cases. An expo- nential decay with respect to frequency refers to chaotic time series while a power-law decay indicates that the spectrum is stochastic. Figure 9 shows a power spectrum obtained by the fast Fourier transform method applied to a telegraph signal mapped from a spike train measured in the red nucleus of a healthy rat (we can use the fast Fourier transform here due to sufficiently large number of spikes in the spike train: 2170). The spike train corresponds to a singular neuron firing. Figure 10 shows analogous spectrum ob- tained from another healthy red nucleus's neuron (2139 spikes). The semi-log scales are used in these figures in order to indicate exponential decay in the spectra (unlike the situation discribed above for a spontaneous activity): in the semi-log scales this decay corresponds to a straight line - Eq. 2. The characteristic frequency fe ≃ 1.1Hz in the both cases. Figure 11 shows a power spectrum ob- tained by the fast Fourier transform method applied to a telegraph signal mapped from a spike train (2022 spikes) measured in the red nucleus of a genetically depressive (the "Flinders" line) rat. The spike train corresponds to a singular neuron firing. Figure 12 shows analogous spectrum obtained from another genetically depressive red nucleus's neuron (2048 spikes). The log-log scales are used in these figures in order to indicate a power- law decay in the spectra (in the log-log scales this decay corresponds to a straight line): E(f ) ∼ f −α (4) In this (scaling) situation there is no characteristic time scale. The scaling exponent α ≃ 1.5 ± 0.1 and ≃ 1.4 ± 0.1 for these two cases. 5 -1.5 -2 -1 0 1 ln f [s-1] 5 4 3 2 1 0 -1 -2 ) f ( E n l FIG. 11: Spectrum of the telegraph signal corresponding to a genetically depressed red nucleus cell. The data are shown in the log-log scales in order to indicate the power law decay Eq. 4 (the straight line). ) f ( E n l 5 4 3 2 1 0 -1 -2 -1.4 -2 -1 0 1 ln f [s-1] FIG. 12: As in Fig. 11 but for another genetically depressed red nucleus cell. V. CHAOTIC NEURAL COHERENCE AND DEPRESSION In order to work together the brain neurons have to make adjustment of their rhythms. The main problem for this adjustment is the very noisy environment of the brain neurons. If their work was based on pure peri- odic inner clocks this adjustment would be impossible due to the noise. The nature, however, has another op- tion. This option is a chaotic clock. In chaotic attrac- tors certain characteristic frequencies can be embedded by broad-band spectra, that makes them much more sta- ble to the noise perturbations [28]. In the light of presented results one can conclude that for the considered cases the healthy neurons firing has de- terministic clocks (periodic and chaotic), while the genet- ically depressive red nucleus's neurons exhibited a pure 1 0.8 0.6 0.4 0.2 y c n e r e h o c healthy depressed 0 0 0.005 0.01 0.02 0.025 0.03 0.015 f [s-1] 6 E(f) ~ exp-f/fe Te=1/fe=0.1s 0 20 40 60 80 100 f [s-1] -140 -150 -160 -170 -180 -190 -200 ] B d [ ) f ( E FIG. 13: Comparison of coherency in firing for the healthy (solid curve) and for the genetically depressed (doted curve) neuron pairs in a low-frequency domain (the in vivo experi- ments). FIG. 14: Spectrum of local field potentials for the posterior piriform (the data were taken from Ref. [34]). The data are shown in the semi-log scales in order to indicate the exponen- tial decay Eq. (2) (the straight line). stochastic firing and it seems that their background de- terministic clocks were broken. The existence of the back- ground clocks can be utilized by the healthy neurons for synchronization of their activity [2],[6],[7],[29]-[31]. In order to compare coherent properties of the healthy and the depressive neuron pairs we will use cross-spectral analysis. The cross spectrum E1,2(f ) of two processes x1(t) and x2(t) is defined by the Fourier transformation of the cross-correlation function normalized by the product of square root of the univariate power spectra E1(f ) and E2(f ): E1,2(f ) = Pτ hx1(t)x2(t − τ )i exp(−i2πf τ ) 2πpE1(f )E2(f ) (5) the bracket h...i denotes the expectation value. The cross spectrum can be decomposed into the phase spectrum φ1,2(f ) and the coherency C1,2(f ): E1,2(f ) = C1,2(f )e−iφ1,2(f ) (6) Because of the normalization of the cross spectrum the coherency is ranging from C1,2(f ) = 0, i.e. no linear relationship between x1(t) and x2(t) at f , to C1,2(f ) = 1, i.e. perfect linear relationship. Figure 13 shows comparison of coherency in firing for the healthy (solid curve) and for the genetically depressed (doted curve) neuron pairs in a low-frequency domain (the in vivo experiments). Despite of the deep anesthesia the healthy neurons exhibit bands of rather high (> 0.5) coherency in the low-frequency domain, while the depres- sive neurons activity is rather decoherent in this domain. VI. LONG-RANGE CHAOTIC COHERENCE The chaotic coherence can involve a large number of the healthy neurons and may be the entire brain. The multi-second oscillations, for instance, are known to be synchronized nearly brain-wide [32],[33]. In the case of depression, however, the chaotic neuron clocks can be broken in a significant part of the brain neurons. That can result in certain decoherence in different parts of the brain. Since depression is usually accompanied by a narrowing of consciousness (and a distorted sense of time) this specific decoherence could be considered as a cause of the phenomenon of the consciousness narrowing as well. The coherence is important for attention, sensorimotor processing, etc.. In humans, in particular, being low in attentional flexibility magnified the effects of private self-focused attention so typical for depressive persons. In order to support the possibility of the extended chaotic coherence we will use analysis of simultaneously recorded local field potentials from three sites along the olfactory-entorhinal axis (the anterior piriform, posterior piriform, and entorhinal cortices: aPIR, pPIR and Ent C) reported in a recent paper [34]. The measurements reported in the Ref. [34] were performed in lightly anesthetized healthy rats (the Long-Evans rats with electrode bundles implanted in their anterior and posterior cortices, and with vertical, silicon probes in their entorhinal cortices), which were emerged from the anesthesia to the waking state with full consciousness. Since the measured local field potentials time series are not spiky ones one does not need in the special data mappings in this case. The authors of the Ref. [34] discovered a new form of coherent neural activity across the three widely separated brain sites, which they named Synchronous Frequency Bursts (SFBs). The high-energy bursts of spontaneous momentary synchrony were observed across widely separated olfactory and entorhinal sites (which have also different architecture: the 6 layers of the entorhinal cortex vs. the three layers of the piriform cortices). Moreover, a significant rate of 0.7 0.6 0.5 0.4 0.3 0.2 0.1 y c n e r e h o c 0 0 5 10 20 25 30 15 f [s-1] FIG. 15: Coherency calculated for the SFBs in the posterior piriform and entorhinal cortices (the data were taken from Ref. [34]). the SFBs simultaneous occurrences was also observed across the different functional processing systems: motor and olfactory ones. The stereotypical duration of the SFBs was about 250 ms and the power spectra taken across the events were exponentially decaying. Figure 14 shows a typical spectrum for the posterior piriform area. The straight line is drawn in this figure in order to indicate the exponential decay Eq. (2) in the semi-log scales (cf. Figs. 9 and 10 for the singular neuron firing). The exponential decay indicates a chaotic nature of these bursts (see above). The decay rate Te = 1/fe ≃ 0.1s is significantly smaller than that observed for the singular neurons (Figs. 9 and 10). Taking into account Eq. (3) one can conclude that the chaotic mixing in the phase space (determined by the Lyapunov's exponents) is much more active for the multineuron activity than for the singular neuron firing (that seems quite natural). This more active mixing shifts the spectrum into more high frequency range. Moreover, one can expect that expansion (globalization) of the chaotic coherence on the larger brain areas will shift the coherent chaotic activity even into the higher frequency ranges (cf. below). The authors of the Ref. [34] computed coherence across SFBs in a pair of brain regions. Figure 15 shows the coherency calculated for the SFBs in the posterior piriform and entorhinal cortices (which are separated in brain space by about 8mm). One can compare this figure with the Fig. 13 (where the coherency was calculated for a pair of neighboring neurons). In this case the fre- quency bands of high coherency can be observed as well. The coherent frequency-range is shifted considerably in the high frequency direction for the multineuron case (see above for a reason of this shift). Actually, "the main purpose of SFBs might be to coordinate multiple frequency bands across different processing subsystems" [34]. Such coordination provides a sufficient level of 7 coherence for the work of these separated subsystems with speed and efficiency impossible in the case of transmission of a specific behavioral content. This can be considered as the main advantage of the chaotic coherence. The hardware for these effective 'manage- ment' can be provided (at least partially) by recently discovered in the cortex and hippocampus interneuronal networks with long-range axonal connections [35],[36] and for the high frequency γ-range (30-90Hz) oscillations "via neurons (and glia) inter-connected by electrical synapses called gap junctions which physically fuse and electrically couple neighboring cells."[37]. The authors of the Ref. [34] observed also that the SFBs occurrence is a function of level of consciousness. They found "that the SFBs occurred far more often under light anesthesia than deeper anesthetic states, and were especially prevalent as the animals regained consciousness". They did not observe the SFBs after the rats regained full alertness, but as they comment this can be a technical problem of inferring the specific signal from the highly complex local field potential of the awake state. Therefore, one cannot rule out the possibility that the phenomenon is still in a full swing also in the fully consciousness state (at least at certain conditions). Finally, it should be noted that the transitional states of consciousness (emerging and decaying) have a very interesting relationship to associative human creativity (H. Poincare called these states as semi-somnolent condi- tions, see Ref. [38], Chapter: Mathematical discovery). The very creative and unexpected associative ideas that come in these states can have the above described long-range chaotic coherence as their direct physical background. Moreover, the same mechanism can also be in work at full consciousness (see previous paragraph). In this case, however, its results are considered as ones coming from the 'clear sky' and we tend to interpret them (may be wrongly) as a result of a prolonged period of unconscious work. In the full consciousness state these results are more often turn out to be adequate ones, unlike of those obtained in the transitional states [38]). "A new result has value, if any, when, by establish- ing connections between elements that are known but until then dispersed and apparently unrelated to one an- other, order is immediately created where chaos seemed to reign" [38]. VII. ACKNOWLEDGMENTS I thank Dremencov E. and Ikegaya Y. for sharing the data and discussions and also Allegrini P. and Grigolini P. for comments and suggestions. I thank Greenberg A. for help in computing. VIII. APPENDIX In order to understand how the fundamental chaotic clock survives the threshold firing let us consider a very simple and rough model, which allows analytic calcula- tion of its autocorrelation function. In this model the spike firing takes place at discrete moments: tn = nT +ζ, where ζ is an uniformly distributed over the interval [0, T ] random variable, n = 1, 2, 3... and T is a fixed period. Then let us consider a telegraph signal constructed for this spike train as it has been described above. If p is a probability of the spike firing at a current moment (0 ≤ p < 1), then the autocorrelation function of such telegraph signal is: C(τ ) = (n − τ /T )(2p − 1)n−1 + (τ /T − (n − 1))(2p − 1)n (A1) in the interval (n − 1)T ≤ τ < nT . Figure 16b shows the autocorrelation function Eq. (A1) calculated for p = 0.25, as an example. For comparison figure 16a shows also a superposition of the autocorrelation functions for the telegraph signals corresponding to the cell-21 (the solid line) and to the spike train generated by the the Rossler attractor fluctuations overcoming the threshold x = 7 (circles). In order to make the autocorrelation functions comparable a rescaling has been made for the Rossler attractor generated autocorrelation function. 8 neuron Rossler a 0 2 4 6 8 10 12 ) τ ( C 1 0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 ) τ ( C 1 0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 τ [s-1] model b T 2T 3T 4T 5T 6T τ FIG. 16: a) Autocorrelation functions for the telegraph sig- nals corresponding to the cell-21 (solid curve) and to the spike train generated by the Rossler attractor fluctuations overcom- ing the threshold x = 7 (circles). b) Autocorrelation function for the simple model telegraph signal: Eq. (A1) with p = 0.25. [1] T. Sauer, Chaos, 5, 127 (1995). [2] E. M. Izhikevich, Dynamical Systems in Neuroscience: The Geometry of Excitability and Bursting (MIT, Cam- bridge, MA, 2006). [3] G. S. Medvedev, Phys. Rev. Lett., 97, 048102 (2006). [4] A. Shilnikov and G. Cymbalyuk, Phys. Rev. Lett., 94, (2002) [15] T. Gedeon , M. Holzer, and M. Pernarowski, Physica D, 178, 267 (2007) [16] N. Crook, W.J. Goh, and M. Hawarat, BioSystems, 87, 267 (2007). [17] T. Pereira , M.S. Baptista, and J. Kurths, Phys. Lett. A, 048101 (2005). 362 159 (2007). [5] H. Korn and P. Faure, C. R. Biologie, 326, 787 (2003). [6] T. Sasaki, N. Matsuki and Y. Ikegaya, J. Neuroscience, [18] O.E. Rossler, Phys. Lett. A, 57 397 (1976). [19] C. Lainscsek, C.Letellier and I. Gorodnitsky, Phys. Lett. 27, 517 (2007). A, 314 409 (2003). [7] N. Takahashi, T. Sasaki, W. Matsumoto, N. Matsuki, and Y. Ikegaya, PNAS, 107, 10244 (2010). [8] A. Bershadskii, E. Dremencov, D. Fukayama and G. Ya- did, Phys. Lett. A, 289, 337 (2001). [9] J.D. Bremner, M. Narayan, E.R Anderson et al., Am. J. Psychiatry, 157, 115 (2000). I. Gorodnitsky dynamics and C. Letellier, [20] C. Lainscsek, from amplitude mea- Reconstructing 8th Joint Symposium sures of spiky time-series, on Neural Computation, (2001) at http://www.its.caltech.edu/∼jsnc/2001/Proceedings/). [21] J.L. Hindmarsh and R.M. Rose, Proc. R. Soc. Lond. B, (available [10] A. Bershadskii and Y. Ikegaya, arXiv:1010.4722 (avail- 221, 87 (1984). able at http://arxiv.org/abs/1010.4722) (2010). [11] N. Ohtomo, K. Tokiwano, Y. Tanaka et. al., J. Phys. Soc. Jpn. 64 1104 (1995). [12] R. Castro and T. Sauer, Phys. Rev. E, 59, 2911 (1999). [13] N. Masuda, and K. Aihara, Phys. Rev. Lett., 88 248101 (2002). [14] K. Aihara and I. Tokuda, Phys. Rev. E, 66, 026212 [22] P. Allegrini, D. Menicucci, R. Bedini, A. Gemignani, and P. Paradisi, Complex intermittency blurred by noise: the- ory and application to neural dynamics Phys. Rev. E, 82, 015103(R) (2010). [23] M. Lukovi´c and P. Grigolini, Power spectra for both in- terrupted and perennial aging processes, J. Chem. Phys. 129, (2008) 184102. 9 [24] P. Grigolini, G. Aquino, M. Bologna, M. Lukovi´c, and B. J. West, A theory of 1/f noise in human cognition, Physica A, 388, (2009) 4192-4204. [25] D.E. Sigeti, Phys. Rev. E, 52, 2443; Physica D, 82, 136 (1995). [26] J. D. Farmer, Physica D, 4, 366 (1982). [27] U. Frisch and R. Morf, Phys. Rev., 23, 2673 (1981). [28] M. I. Rabinovich, and H. D. I. Abarbanel, Neuroscience, 061904 (2005). [32] D. Contreras, and M. Steriade, J. Neurosci., 15, 604 (1995). [33] M. Steriade, C Neuroscience 101 243 (2000). [34] R. Hermer-Vazquez, L. Hermer-Vazquez, and S. Srini- vasan, Brain Res. Bull., 79, 6(2009). [35] T. Klausberger, P. Somogyi, Science, 321 53 (2008). [36] S. Jinno, T. Klausberger, L.F. Marton, et al., J. Neurosci. 87, 5 (1998). 27 8790 (2007). [29] D.I. Abarbanel, R. Huerta, M. I. Rabinovich, et al.,, Neu- ral Comput., 8, 1567 (1996). [37] S. Hameroff, J. Biol. Phys., 36, 71-93 (2010). [38] H. Poincare, Science and Method (Courier Dover Publi- [30] F. Mormann, K. Lehnertz, P. David, et al., Physica D, cations, NY, 2003). 144, 358 (2000). [31] E. Rossoni, Y. Chen, M. Ding, et al., Phys. Rev. E, 71,
1803.08109
4
1803
2018-08-09T15:56:54
A mechanistic model of connector hubs, modularity, and cognition
[ "q-bio.NC", "q-bio.QM" ]
The human brain network is modular--comprised of communities of tightly interconnected nodes. This network contains local hubs, which have many connections within their own communities, and connector hubs, which have connections diversely distributed across communities. A mechanistic understanding of these hubs and how they support cognition has not been demonstrated. Here, we leveraged individual differences in hub connectivity and cognition. We show that a model of hub connectivity accurately predicts the cognitive performance of 476 individuals in four distinct tasks. Moreover, there is a general optimal network structure for cognitive performance--individuals with diversely connected hubs and consequent modular brain networks exhibit increased cognitive performance, regardless of the task. Critically, we find evidence consistent with a mechanistic model in which connector hubs tune the connectivity of their neighbors to be more modular while allowing for task appropriate information integration across communities, which increases global modularity and cognitive performance.
q-bio.NC
q-bio
A mechanistic model of connector hubs, modularity, and cognition Maxwell A. Bertolero1,2*, B.T. Thomas Yeo3,4, Danielle S. Bassett2,5-7, Mark D'Esposito1 1Helen Wills Neuroscience Institute and the Department of Psychology, University of California, Berkeley, CA 94720-3190, USA 2Department of Bioengineering, School of Engineering and Applied Sciences, University of Pennsylvania, Philadelphia, PA 19104 USA 3Electrical and Computer Engineering, ASTAR-NUS Clinical Imaging Research Centre, Singapore Institute for Neurotechnology & Memory Networks Programme, National University of Singapore, Singapore 119077, Singapore 4NUS Graduate School for Integrative Sciences and Engineering, National University of Singapore, Singapore 5Department of Electrical & Systems Engineering, School of Engineering and Applied Sciences, University of Pennsylvania, Philadelphia, PA 19104, USA 6Department of Physics & Astronomy, College of Arts and Sciences, University of Pennsylvania, Philadelphia, PA 19104 USA 7Department of Neurology, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104 USA. *corresponding author ([email protected]) Abstract The human brain network is modular -- comprised of communities of tightly interconnected nodes1. This network contains local hubs, which have many connections within their own communities, and connector hubs, which have connections diversely distributed across communities2,3. A mechanistic understanding of these hubs and how they support cognition has not been demonstrated. Here, we leveraged individual differences in hub connectivity and cognition. We show that a model of hub connectivity accurately predicts the cognitive performance of 476 individuals in four distinct tasks. Moreover, there is a general optimal network structure for cognitive performance -- individuals with diversely connected hubs and consequent modular brain networks exhibit increased cognitive performance, regardless of the task. Critically, we find evidence consistent with a mechanistic model in which connector hubs tune the connectivity of their neighbors to be more modular while allowing for task appropriate information integration across communities, which increases global modularity and cognitive performance. Main The human brain is a complex network that can be parsimoniously summarized by a set of nodes representing brain regions and a set of edges representing the connections between brain regions. In network models of fMRI data, each edge represents the strength of functional connectivity -- the temporal correlation of fMRI activity levels -- between the two nodes (Figure 1). This network model can be used to study global and local brain connectivity patterns. Brain networks contain communities -- groups of nodes that are more strongly connected to members of their own group than to members of other groups(Figure 1)1,4,5. This feature of networks is termed modularity and can be quantified by the modularity quality index Q (see Methods for equation). Modularity is ubiquitously observed in complex systems in Nature -- a modular structure is observed consistently across the brains of very different species, from c elegans to humans6. Given its ubiquity, modular network organizations are potentially naturally selected because they reduce metabolic costs. Functional and structural connectivity is metabolically expensive7-11. A modular architecture with anatomically segregated and functionally specialized communities reduces the average length and number of connections -- the network's wiring cost. Moreover, the brain's genotype-phenotype map is modular, forming groups of phenotypes, including brain communities (e.g., the visual community)12,13, that are co-affected by groups of genes14. Modularity, at the genetic and phenotypic level, allows systems to quickly evolve under new selection pressures15,16. As we noted in earlier work2, modularity potentially increases fitness in information processing systems 17-19, and network simulations show that modularity allows for robust network dynamics, in that the connections between nodes can be reconfigured without sacrificing information processing functions, a process necessary for the evolution of a network20. Artificial intelligence research has shown that modular networks also solve tasks faster and more accurately and evolve faster than non-modular networks 21 with lower wiring costs than non-modular networks22. Critically, modularity is also behaviorally relevant -- modularity predicts intra-individual variation in working memory capacity23 and how well an individual will respond to cognitive training24,25. Within each of these communities, local hubs exist that have strong connectivity to their own community. The within community strength can be used to measure a node's locality (see Methods for equation; Figure 1 for a schematic). A high within community strength reflects that a node has strong connectivity within its own community and is thus a local hub. Local hubs are ideally wired for segregated processing. Because the connections of local hubs are predominantly concentrated within their own community and their functions are likely specialized and segregated, damage to local hubs tends to cause relatively specific cognitive deficits26,27 and does not significantly alter the modular organization of the network27. Supporting their more segregated and discrete role in information processing, their activity levels do not increase as more communities are involved in a task1. Yet, a completely modular organization renders the brain extremely limited in function -- without connectivity between communities, information from, for example, visual cortex could never reach motor cortex and therefore visual information could not be used to inform movements. Thus, how is information integrated across these mostly segregated communities? The interdependence between modular communities and integration is a modern rendition of one of the first observations in neuroscience -- Cajal's conservation principle, which states that the brain has been naturally selected and is thus organized by an economic trade-off between minimizing the wiring cost of the network, which leads to modularity, and more costly connectivity patterns that increase fitness, like the integrative functions afforded by connections between communities 28-30. Connector hubs have diverse connectivity across different communities. The participation coefficient can be used to measure a node' diversity (see Methods for equation, Figure 1 for a schematic). A high participation coefficient reflects that a node has connections equally distributed across the brain's communities and is thus labeled a connector hub. Connector hubs are ideally wired for integrative processing1,28,31-34. In human brain networks, connector hubs have a particular cytoarchitecture35, are implicated in a diverse range of cognitive tasks36,37, and are physically located in anatomical areas at the boundaries between many communities33. Moreover, damage to connector hubs causes widespread cognitive deficits26 and a decrease in the modular structure of the network27. During cognitive tasks, connector hubs appear to coordinate connectivity changes between other pairs of nodes -- activity in connector hubs predicts changes in the connectivity of other nodes, particularly the connectivity between nodes in different communities38-40. Connector hubs are also strongly interconnected to each other, forming a diverse club -- tightly interconnected connector hubs5. Connector hubs also have connections to almost every community in the network. Thus, they have access to information from every community. Finally, connector hubs exhibit increased activity if more communities are engaged in a task, which suggests that connector hubs are involved in processes that are more demanding as more communities are engaged1. Connector hubs might be Nature's cheapest solution to integration in a modular network. Generative models suggest that the diverse club -- tightly interconnected connector hubs -- potentially evolved to balance modularity and efficient integration5. However, given the amount of wiring required to link to many different and distant communities, connector hubs' connectivity pattern dramatically increases wiring costs11. Despite this cost, connector hubs potentially provide a necessary function -- connector hubs could be the conductor of the brain's neural symphony. A parsimonious mechanistic model of these findings is that connector hubs tune connectivity between communities. Neuronal tuning refers to cells selectively representing a particular stimulus, association, or information. We introduce the mechanistic concept of network tuning, in which connections between nodes are organized to achieve a particular network function or topology, like the integration of information across communities or decreased connectivity between two communities. We propose that diverse connectivity across the network's communities allows connector hubs to tune connectivity between communities to be modular but also allows for task appropriate information integration across communities. This facilitates a global modular network structure in which local hubs and nodes within each community are dedicated to mostly autonomous local processing. The modular network structure afforded by diversely connected connector hubs -- connector hubs that are wired well for network tuning -- is potentially optimal for a wide variety of cognitive processes. Thus, despite their cost, strong and diverse connector hubs might be critically necessary for integrative processing in complex modular neural networks. Local and connector hubs have been exhaustively studied by network science and their functions have been inferred from their topological locations in the network5. Moreover, individuals' brain network connectivity has been shown to be predictive of task performance41-45,46 and is able to "fingerprint" individuals47. However, no study has leveraged these individual differences to test a mechanistic model of hub function and direct evidence supporting a mechanistic model of these hubs and how they support human cognition remains absent. Moreover, it is currently unknown if there is a hub and network structure that is optimal for a diverse set of tasks or if different hub and network structures are optimal for different tasks. Here, we analyze how individual differences in the locality and diversity of hubs during the performance a task relates to network connectivity, modularity, and performance on that task as well as subject measures collected outside of the scanner, including psychometrics (e.g., fluid intelligence, working memory) and other behavioral measures (e.g., sleep quality and emotional states). We test a mechanistic model in which connector hubs tune their neighbors' connectivity to be more modular, which increases the global modular structure of the network and task performance, regardless of the particular task. We leveraged the size and richness of fMRI data from 476 (S500 release) subjects that participated in the Human Connectome Project48. A network was built for each subject using fMRI data collected during seven different cognitive states (Resting-State, Working Memory, Social Cognition, Language & Math, Gambling, Relational, Motor, see Methods). Thus, each subject has seven different networks for analysis. Each edge represents the strength of functional connectivity between each pair of 264 nodes49. In every network, Q and a division of nodes into communities was calculated (see Methods). Next, in every network, for each node, locality and diversity was measured. Within community strength measures a node's locality and the participation coefficient measures a node's diversity, respectively (see Methods for equations). Figure 1 displays this processing workflow. Figure 1 Functional connectivity and network science processing workflow. a, The mean signal across time is extracted from 264 cortical, sub-cortical, and cerebellar regions, three of which are shown here. b, The time series of the three nodes is shown. To measure functional connectivity, the Pearson r correlation coefficient between the time series of node i and the time series of node j for all i and j is calculated. c, The strongest (e.g., the top 5% percent r values) functional connections serve as weighted edges in the graph (a range of graph densities was explored, see Methods for details). d, The Infomap community detection algorithm is applied, generating a community assignment for each node, displayed here in different colors in a schematic (top) and the mean graph across subjects (bottom). e, Given that particular community assignment and network, nodes' participation coefficients are calculated. Red nodes are high participation coefficient nodes, shown here in a schematic (top) and the mean graph (bottom). f, Within community strengths are also calculated. Purple nodes are high within community strength nodes, shown here in a schematic (top) and the mean graph (bottom). The graphs along the bottom are laid out using the force-atlas algorithm, where nodes are repelling magnets and edges are springs, which causes nodes in the same community to cluster together, nodes that are diversely connected across communities (connector hubs) to be in the center of the graph, and nodes that are strongly connected to a single community (local hubs) in the middle of that community. d, lower, A single community (light blue) and its connections to the rest of the graph is extracted and enlarged, with nodes colored by community. Note that the nodes within each community are more strongly connected to each other than to nodes in other communities. e, lower, A node (and its connections) with a high participation coefficient is extracted and enlarged, with nodes colored by community. Note that the connector hub is connected to many different communities. f, A node (and its connections) with a high within community strength is extracted and enlarged, with nodes colored by community. Note that the local hub is strongly connected to its own community. In the proposed mechanistic model of hub connectivity, connector hubs, via their diverse connectivity, tune the network to preserve or increase global modularity and local hubs' locality, which, in turn, increases task performance. If this model is true, hubs' connectivity in the network and network modularity should be predictive of task performance. Thus, the first test of this model involved using hub diversity and locality, network connectivity, and modularity to predict task performance. A predictive multilayer perceptron model (three layers plus the input layer and the output layer (enough for non-linear relationships); eight neurons per layer (one per feature, with two layers containing 12 neurons, allowing for higher dimensional expansion)) was used to predict subjects' task performance (Supplementary Figure 1, Figure 2). Known as deep neural networks, these predictive models are constructed by tuning the weights between neurons across adjacent layers to achieve the most accurate relationship between the features (input) and the value the model is trying to predict (output). The predictive model's features (n=8) captured how well subjects' nodes' diversity and locality, network connectivity (i.e., edge weights in the network), and modularity (Q) are optimized for the performance of a task. For example, for the feature that captures how optimized the diversity of subjects' nodes' are for task performance, for each node, the Pearson r across subjects between that node's participation coefficients (which measures diversity) and task performance values was calculated (Supplementary Figure 1a). We call this r value the node's diversity facilitated performance coefficient. The feature, then, for a given subject, is the Pearson r across nodes between each node's diversity facilitated performance coefficient and each node's participation coefficient in that subject, representing how optimized the diversity of that subject's nodes' are for performance in the task (Supplementary Figure 1). Critically, for each subject's feature calculation, the diversity facilitated performance coefficients are calculated without that subject's data. The same procedure is executed for locality (using the within community strengths) and edge weights; instead of participation coefficients, within community strengths or edge weights are used. Finally, the Q values of the networks are included in the model. The predictive model was fit for each of the four cognitive tasks that subjects performed in the Human Connectome Project for which performance was measured (Working Memory, Relational, Language and Math, Social tasks; see Methods for task performance measures). For this and other subject performance analyses, we could not analyze the Gambling, Motor, or Resting-State tasks, as there was no performance measured for these tasks. Each predictive model was fit to the subjects' networks constructed during the performance of each task as well as the resting state (four features from each). The inclusion of the resting-state and the cognitive task state allowed the model to capture the subjects' so-called intrinsic network states as well as the subjects' task driven network states. Using a leave-one-out cross-validation procedure, the features were constructed, and the model was fit with data from all subjects except one. The predictive model was then used to predict the left-out subject's task performance (Supplementary Figure 1c). To test the accuracy of the model, the Pearson r between the observed and predictive performance of each subject was calculated (Figure 2, Supplementary Figure 2). This predictive model significantly (p<0.001, Bonferroni corrected (n tests = 4)) predicted performance in all four tasks (Figure 2). Also, using a predictive model with only nodes' diversity and locality and modularity features (i.e., ignoring individual connections in the network) did not dramatically decrease the models' prediction accuracies (Supplementary Figure 2a,b). Given that head motion is a concern when analyzing fMRI data, scrubbing techniques were applied to remove motion artifacts from the fMRI data and the mean frame-wise displacement was regressed out from task performance. Neither of these additional analyses dramatically decreased the predictive models' prediction accuracies (Supplementary Figure 2c-f). Finally, in each task, modularity (Q) alone was only weakly correlated with task performance (Working Memory, Pearson's r (dof=471):0.303, p<0.001, CI:0.219,0.383; Relational, Pearson's r (dof=455):0.106, p:0.095, CI:0.014,0.196; Language & Math, Pearson's r (dof=469):0.085, p:0.259, CI:-0.005,0.174; Social, Pearson's r (dof=471):0.084, p:0.275, CI:-0.006,0.173, all confidence intervals=95%). These results suggest that the diversity and locality of nodes, in combination with the modular connectivity structure of the network, are highly predictive of task performance. The Human Connectome project contains psychometrics and other behavioral measures collected outside of the MRI scanner; for clarity and to differentiate these measures from the task performance measures and the tasks' corresponding networks, we call these "subject measures" 50. If a particular hub and network structure is generally optimal for many different types of cognition and many different behaviors (a component of the mechanistic model of hub function), then the tasks' optimal hub and network structures should be similarly optimal across subject measures -- sub-optimal for negative measures like poor sleep, sadness, and anger and optimal for positive measures like life satisfaction and processing speed. Figure 2 Hub diversity and locality, modularity, and network connectivity predict cognitive performance. a, for each task, the correlation between task performance and the performance predicted by a predictive model of hub diversity and locality, modularity, and network connectivity. Each data point represents the (y-axis) true performance (see Methods, each task's performance value scale is unique) of the subject and the (x-axis) predicted performance of the subject by the neural network. Shaded areas represent 95 percent confidence intervals. In every task, the predictive model significantly predicted task performance (p < 1e-3, Bonferroni corrected (n tests=4), N=Working Memory: 473, Relational: 457, Language & Math: 471, Social: 473). b, we correlated the tasks' feature correspondence values (see Supplementary Figure 3 for each task's feature correspondence with each subject measure) -- measuring if the two tasks' optimal hub and network structures are also optimal for the same subject measures. High correlations mean that the two tasks' hub and network structures are similarly optimal for the same subject measures (all results significant at p < 1e-3, Bonferroni corrected (n tests = 4), dof=45, N=47, the number of subject measures, while the feature correspondence N =Working Memory: 473, Relational: 457, Language & Math: 471, Social: 473). For each task, the predictive model constructs features that capture how optimal each subject's hub and network structure is for performance on that task. Using the subjects' networks from a given task, the predictive model of hub and network structure can construct features that capture how optimal each subject's hub and network structure is for a given subject measure collected outside of the scanner instead of task performance. This was executed using the networks from each of the four tasks for all subject measures. Thus, the predictive model constructs features that capture how optimal each subject's hub and network structure (measured during the performance of a task (e.g. Working Memory)) is for a given subject measure (e.g., Delayed Discounting). The correspondence between the features in the two models -- how similarly optimal subjects' hub and network structure are for the task and a given subject measure -- can then by calculated by, across subjects, computing the correlation between the features in the two predictive models. Specifically, for each feature, the correlation, across subjects, between the feature in the predictive model fit to task performance (e.g., Working Memory) and that feature in the predictive model that was fit to a subject measure (e.g., Delayed Discounting) is computed. The mean correlation across the edge, locality, and diversity features (n=6, three features from resting-state and three features from the task) is then calculated, which we call feature correspondence. The Q feature was ignored, as the Q feature remains constant regardless of what the model is fit to. Thus, this value determines if each task's optimal hub and network structure is optimal for other subject measures and if all the tasks' optimal hub and network structures are similarly optimal for other subject measures. For each task, the hub and network structures that were optimal for that task were typically also optimal for positive subject measures and sub-optimal for negative subject measures (Supplementary Figure 3). Next, the similarity by which two tasks' optimal hub and network structures generalized to other subject measures can be measured by correlating the feature correspondence values (for example, the Working Memory and Social columns in Supplementary Figure 3). High Pearson r correlations were found between all tasks (r (dof=45) values between 0.82 and 0.96, p<0.001 Bonferroni corrected (n tests = 4), Figure 2b). Finally, the predictive model was able to significantly predict most subject measures (Supplementary Figure 4). These results demonstrate that, if an individual has a particular brain state during a given task, as defined by the connectivity of the network's hubs, that is optimal for that given task, it also likely optimal for other subject measures. Critically, different tasks' optimal hub and network structures are similarly optimal for other subject measures. Moreover, these findings demonstrate that the predictive model captures hub connectivity patterns in the network that are relevant for behavior and cognition in general, instead of overfitting hub connectivity patterns that are only related to a particular cognitive process or behavior. Having established relationships between hub locality and diversity, modularity, and task performance, we sought to test the mechanistic claim that diverse connector hubs increase modularity by analyzing how individual differences in a node's diversity within the network are predictive of individual differences in brain network modularity (Q; see Methods for mathematical definition). Typically, the result of damage to a region can be used to infer the function of that region -- if a region is damaged and modularity decreases, the region is putatively involved in preserving modularity. Here, we analyze the other direction -- when a hub is diversely connected across the brain (i.e., strong), if modularity increases, the region's diverse connectivity is putatively involved in preserving modularity (Supplementary Figure 5). Thus, we first tested if, across subjects, a node's participation coefficients are positively correlated with modularity (Q). For each node, the Pearson r between that node's participation coefficients and the network's modularity values (Q) across subjects was calculated. Intuitively, higher r values indicate that the node's diversity (i.e., the participation coefficient) is associated with higher network modularity. This is an important feature that can be used to distinguish the roles of different brain regions. For ease of presentation, we refer to each node's r value as the diversity facilitated modularity coefficient, as it measures how the diversity of the node's connections facilitates (we use this term to remain causally agnostic) the modularity of the network. For every node, the Pearson r between the within community strengths and Q values across subjects was also calculated. Intuitively, higher r values indicate that the node's locality (i.e., the within community strength) is associated with higher network modularity. We refer to each node's r value (between within community strengths and Q values across subjects) as the locality facilitated modularity coefficient, as it measures how the locality of the node's connections facilitates the modularity of the network. We performed these computations separately for all seven distinct cognitive states. In all states, the diversity facilitated modularity coefficients of connector hubs (top 20 percent highest participation coefficient nodes) were shown to be significantly higher than other nodes in a Bonferroni-corrected independent two-tailed t-test (Figure 3a, Working Memory t(dof:262):7.182, p<0.001, Cohen's d:1.104, CI:0.062,0.117, Gambling t(dof:262):4.101, p:0.0004, Cohen's d:0.63, CI:0.025,0.052, Language & Math t(dof:262):7.292, p<0.001, Cohen's d:1.12, CI:0.062,0.102, Motor t(dof:262):7.354, p<0.001, Cohen's d:1.13, CI:0.088,0.13, Relational t(dof:262):4.457, p:0.0001, Cohen's d:0.685, CI:0.038,0.075, Resting State t(dof:262):3.947, p:0.0007, Cohen's d:0.606, CI:0.029,0.096, Social t(dof:262):3.716, p:0.0017, Cohen's d:0.571, CI:0.022,0.051. P values Bonferroni corrected (n tests=7), all confidence intervals=95%). Moreover, in all cognitive states, the locality facilitated modularity coefficients of local hubs (top 20 percent highest within community strength nodes) were shown to be significantly higher than other nodes in a Bonferroni-corrected independent two-tailed t-test (Figure 3b, Working Memory t(dof:262):5.415, p<0.001, Cohen's d:0.832, CI:0.045,0.093, Gambling t(dof:262):4.959, p<0.001, Cohen's d:0.762, CI:0.034,0.074, Language & Math t(dof:262):6.428, p<0.001, Cohen's d:0.988, CI:0.045,0.085, Motor t(dof:262):9.822, p<0.001, Cohen's d:1.509, CI:0.101,0.146, Relational t(dof:262):6.131, p<0.001, Cohen's d:0.942, CI:0.036,0.07, Resting State t(dof:262):0.966, p:1.0, Cohen's d:0.148, CI:-0.014,0.038, Social t(dof:262):4.54, p:0.0001, Cohen's d:0.698, CI:0.026,0.06. P values Bonferroni corrected (n tests = 7), all confidence intervals=95%). While the diversity facilitated modularity coefficients of connector hubs were not always positive, they were typically close to or above zero. This means that a diverse connector hub can be associated with increased integrative connectivity between communities without decreasing the modularity of the network. To more fully understand the relationship between nodes' diversity and the networks' modularity, the Pearson r between each node's mean participation coefficient across subjects (which defines a connector hub) and the node's diversity facilitated modularity coefficient was calculated (Supplementary Figure 6). Moreover, the Pearson r between each node's mean within community strength across subjects (which defines a local hub) and the node's locality facilitated modularity coefficient was calculated (Supplementary Figure 6). In every task, there was a significant positive correlation between a node's mean participation coefficient and that node's diversity facilitated modularity coefficient (Supplementary Figure 6a). In every task, there was also a significant positive correlation between a node's mean within community strength and its locality facilitated modularity coefficient (Supplementary Figure 6b). These analyses demonstrate that connector hubs' strong diverse connectivity to many communities and local hubs' strong local connectivity is associated with higher brain network modularity, regardless of the subjects' cognitive state. Thus, these results are consistent with the mechanistic model of connector hub function, where connector hubs preserve the modular structure of the network via diverse connectivity. We confirmed the reliability and reproducibility of these results and demonstrated that they are not driven by analytically necessary relationships. First, the mean participation coefficient and within community strength was calculated in one half of the subjects and the diversity and locality facilitated modularity coefficients were calculated in the other half of the subjects, testing 10,000 splits (Supplementary Figure 7). Next, four null models were tested to ensure the current results were not driven by analytically necessary relationships (Supplementary Figure 8). Other analyses ensured the current results are not driven by the number of communities (Supplementary Figure 9, Supplementary Figure 10). Finally, to justify the use of the Pearson r to calculate the coefficients, the relationship between nodes' diversity and Q was confirmed as typically linear (Supplementary Figure 11). Having found evidence supporting a mechanistic model in which connector hubs tune their neighbors' connectivity to be more modular, thereby increasing the global modular structure of the network, we next asked if diverse hubs concurrently facilitate higher modularity and higher task performance. To address this question, the Pearson r between each node's participation coefficient (or within community strength) and task performance was calculated. A positive r at a node indicates that a subject with a higher participation coefficient (or within community strength) at that node performs better on the task. We refer this r value as the node's diversity facilitated performance coefficient for participation coefficients, and locality facilitated performance coefficient for within community strengths. Note that these are the same r values used in the construction of the predictive performance model features (Supplementary Figure 1). However, for the predictive model, we calculated these r values with the subject whose behavior was to be predicted held out. Here, we calculate these r values across all subjects. In all tasks, the diversity facilitated performance coefficients of connector hubs (top 20 percent strongest) were shown to be significantly higher than other nodes in a Bonferroni-corrected independent two-tailed t-test, with only the Language & Math task at p=0.0677 after Bonferroni correction (uncorrected p=0.0169) (Figure 3c, Working Memory t(dof:262):5.378, p<0.001, Cohen's d:0.826, CI:0.03,0.071, Language & Math t(dof:262):2.404, p:0.0677, Cohen's d:0.369, CI:0.005,0.04; Relational t(dof:262):2.959, p:0.0135, Cohen's d:0.455, CI:0.01,0.037; Social t(dof:262):4.744, p<0.001, Cohen's d:0.729, CI:0.025,0.053. All p values Bonferroni corrected (n tests=4), all confidence intervals=95%). The locality facilitated performance coefficients of local hubs (top 20 percent strongest) were shown to be significantly higher than other nodes in a Bonferroni-corrected independent two-tailed t-test (Figure 3d, Working Memory t(dof:262):2.712, p:0.0285, Cohen's d:0.417, CI:0.008,0.054, Language & Math t(dof:262):2.864, p:0.0181, Cohen's d:0.44, CI:0.006,0.043; Relational t(dof:262):0.327, p:1.0, Cohen's d:0.05, CI:-0.016,0.021; Social t(dof:262):1.862, p:0.2547, Cohen's d:0.286, CI:0.0,0.031. All p values Bonferroni corrected (n tests=4), all confidence intervals=95%). Moreover, the correlation between each node's diversity facilitated performance coefficient and each node's mean participation coefficient was positive and significant (Supplementary Figure 6), suggesting that, for connector hubs, a higher participation coefficient is associated with higher task performance. The Pearson r correlation between each node's locality facilitated performance coefficient and each node's mean within community strength was also positive and significant (Supplementary Figure 6; all tasks except Relational, Bonferroni (n tests=4) p=0.081, uncorrected p=0.02), suggesting that, for local hubs, a higher within community strength is also associated with higher task performance. Finally, there was a significant positive correlation between a node's diversity facilitated modularity coefficient and a node's diversity facilitated performance coefficient (Figure 3e) as well as a significant positive correlation between a node's locality facilitated modularity coefficient and a node's locality facilitated performance coefficient (Figure 3f). Thus, diverse connector hubs facilitate higher task performance in proportion to how much they facilitate higher modularity, suggesting a strong link between the increased modularity afforded by diverse hubs and increased task performance. Figure 3 Connector hubs and local hubs concurrently facilitate increased modularity and task performance. For each task, diversity and locality facilitated modularity coefficients, a measure of how the diversity and locality (respectively) of a node facilitates modularity, were calculated. In every task, the diversity and locality facilitated modularity coefficients of connector (a) and local hubs (b), compared to other nodes, is significantly (except Resting-State for locality) higher, demonstrating that strong connector and local hubs facilitate the modular structure of brain networks. For each task, diversity and locality facilitated performance coefficients were calculated. In every task the diversity and locality facilitated performance coefficients of connector (c) and local hubs (d), compared to other nodes, is significantly (except Language for diversity (p=0.0677 after Bonferroni correction (uncorrected p=0.0169)), Relational and Social for locality) higher, demonstrating that strong connector and local hubs facilitate increased task performance. For a-d, the mean and quartiles are marked in each violin. Each task's distribution of coefficients was tested for normality using D'Agostino and Pearson's omnibus test k2. No evidence was found (k2>0.0 for all tasks) that these distributions were not normal. N=264, the number of nodes in the graph. e, The correlation between a node's diversity facilitated modularity coefficient and a node's diversity facilitated performance coefficient. f, The correlation between the node's locality facilitated modularity coefficient and the node's locality facilitated performance coefficient. In panels e,f, N=264, the number of nodes in the graph. Shaded areas represent 95 percent confidence intervals. All p values are Bonferroni corrected (n tests = 4). Next, we tested the mechanistic network tuning claim of the model: "do connector hubs increase Q by tuning the connectivity of their neighbors' edges to be more modular?" This relationship should only hold for connector hubs, not local hubs, as previous studies suggest that connector hubs tune connectivity between communities and maintain a modular structure1,36,38-40. We therefore examined connector hubs for which their diversity facilitated modularity coefficients were positive. This analysis had two aspects. First, do connector hubs increase modularity by tuning within community edge strengths? Second, are connector hubs tuning the within community edge strengths of their neighbors in order to increase global modularity? In order to test the first aspect of the neural tuning mechanism -- if within community edges are tuned by connector hubs in order to increase global modularity -- we used a canonical division of nodes into communities (Figure 4d displays this division, see Methods for link to division36,49). We assessed, for each edge in the network, how the edge's weights related to modularity values (Q) across subjects (Figure 4a). Next, we calculated how well each connector hub's participation coefficients correlate with each edge's weights across subjects (Figure 4b). Higher Q values and higher connector hub participation coefficients are associated with decreased connectivity between the visual, sensory/motor hand, sensory/motor mouth, auditory, ventral attention, dorsal attention, and cingulo-opercular communities. These communities were also more strongly connected to fronto-parietal, default mode, salience, and sub-cortical communities in networks with higher modularity values and higher connector hub participation coefficients. Given these observations, we sought to find the edges that mediate between connector hubs' increased participation coefficients and modularity (Q), as these are the edges that connector hubs likely tune in order to increase Q. Specifically, a mediation analysis was performed for each connector hub, with an edge weight mediating the relationship between the connector hub's participation coefficients and the Q indices of the networks across subjects. An edge's mediation value of a connector hub's participation coefficients and Q is the regression coefficient of the edge's weights by the connector hub's participation coefficients across subjects multiplied by the regression coefficient of Q indices by the edge's weights, controlling for the connector hub's participation coefficients, across subjects. Each edge's mean mediation value across connector hubs is shown in Figure 4c. We found that edges between the visual, sensory/motor hand, sensory/motor mouth, auditory, ventral attention, dorsal attention, and cingulo-opercular communities, as well as edges between those communities and the fronto-parietal, default mode, and sub-cortical communities, mediate the relationship between connector hubs' participation coefficients and Q indices. These results are consistent with a mechanistic model in which diverse connector hubs tune connectivity to increase segregation between sensory, motor, and attention systems, which increases the global modularity of the network. Next, we tested the second aspect of the network tuning mechanism -- if the relationship between a connector hub's participation coefficients and Q indices is mediated primarily by that connector hub's neighbors' edge pattern increasing Q. Neighbors were defined based on edges present between the two nodes in a graph at a density of 0.15 (as it was our densest cost explored). The mediation values calculated above each represent an edge mediating between a node i's participation coefficients and Q values. Thus, for each connector hub i, there is the set of arrays of absolute mediation values of node i's neighbors' edges (n=263 for each neighbor j's array) and the set of arrays of the absolute mediation values of node i's non-neighbors' edges (n=263 for each non- neighbor j's array). Edges of node i in every array were ignored, as we were only interested in how the participation coefficients of connector hub i modulate Q via the mediation of j's connectivity to the rest of the network, not j's connectivity to connector hub i (thus, n=264-1). If a connector hub is primarily modulating Q via the tuning of its neighbors' edges, then the absolute mediation values in the neighbors' arrays should be greater than the absolute values in the non-neighbors' arrays. The distribution of t- values between the two sets of arrays for all connector hubs (neighbors versus non- neighbors) is shown in Supplementary Figure 12; across tasks, the mediation values were consistently and significantly higher for connector hubs' neighbors' edges than non-neighbors edges. Moreover, these same t-values can be calculated for local hubs, using the within community strength instead of the participation coefficient; thus, an edge mediates between a local hub's within community strengths and Q. Across tasks, connector hubs' neighbors' mediation t values were shown to be higher than local hubs' neighbors' mediation t values with a two-tailed independent student's t-test (t(dof:1358): 3.892, p:0.0001, Cohen's d:0.219, CI:1.887,6.62), demonstrating that this result is specific to connector hubs. All distributions were confirmed as normal (k2>100.0, p<0.00001 for all tasks). We also performed an alternative analysis that confirmed these relationships (see Methods, Supplementary Figure 13). These results suggest that each connector hub, not local hub, tunes their neighbors' connectivity to be more modular. A connector hub's high diversity facilitated modularity coefficient does not largely reflect diffuse global connectivity changes. Instead, connector hubs are likely connected in a way that allows them to directly tune the connectivity of their neighbors to be more modular, thereby increasing global network modularity. Thus, the locality facilitated modularity coefficients are likely a downstream effect of connector hub modulation. Supporting this interpretation, we found that, when connector hubs have high participation coefficients, local hubs have high within community strengths (Supplementary Figure 14). Figure 4 Connectivity between primary sensory, motor, dorsal attention, ventral attention, and cingulo-opercular communities mediate the relationship between connector hubs and modularity. a, Each entry is the Pearson correlation coefficient, r, across subjects (N=476), between modularity (Q) and that edge's weights. b, For each connector hub, the Pearson r between the hub's participation coefficients and each edge's weights across subjects (N=476) was calculated. The matrix in b is the mean of those matrices across connector hubs. c, To investigate the relationship between connector hubs' participation coefficients, edge weights, and Q, a mediation analysis was performed for each connector hub, with an edge's weights mediating the relationship between the connector hub's participation coefficients and Q indices (N=476). Each edge's mean mediation value between connector hubs' participation coefficients and Q is shown. d, The anatomical locations of each node and community on the cortical surface36,49. In the series of analyses we report here, we explicitly and comprehensively tested a mechanistic model by leveraging individual differences in connectivity and cognition in humans. Specifically, a model of the diversity and locality of hubs, the modularity of the network, and the network's connectivity was highly predictive of task performance and a range of subject measures. Critically, the diversity and locality of nodes optimal for each task were also similarly optimal for positive subject measures. Thus, it appears that there is a hub and network structure that is generally optimal for cognitive processing. We found evidence that diverse connector hubs preserve or increase the modularity of brain networks. Moreover, diverse connector hubs tune the connectivity of their neighbors to be more modular. Finally, we found that the diversity of connector hubs simultaneously facilitated higher modularity and task performance. Thus, connector hubs appear to contribute to the maintenance of an optimal modular architecture during integrative cognition without greatly increasing the wiring cost or decreasing modularity5,29,51. In sum, these data are consistent with a mechanistic model of hub function, where connector hubs integrate information and subsequently tune their neighbors' connectivity to be more modular, which increases the global modularity of the network, allowing local hubs and nodes to perform segregated processing. Across individuals, we found that diverse connector hubs increase modularity and task performance, regardless of the task. In all seven tasks, the subjects with the most diversely connected connector hubs also had the highest modularity and, in the four tasks for which performance was measured, the highest task performance. Thus, we propose that connector hubs are likely critical for integrating information and tuning their neighbors' connectivity to be more modular, regardless of the task. Although connector hubs are more active during tasks that require many communities1, as their functions are likely more computationally demanding during these tasks, it is likely that every task requires the functions of connector hubs, as supported by our finding that their diverse connectivity predicts performance in all of the tasks analyzed here. Our findings compliment many previous task-based fMRI studies that have identified regions that are more active during a particular cognitive process. We have demonstrated that, while different regions are more or less active in different tasks, including connector hubs, the diverse connectivity and integration and tuning functions of connector hubs are consistently required across different cognitive processes. Our findings are also consistent with neuropsychological studies of patients with focal brain lesions. It has been found that damage to connector hubs decreases modularity and causes widespread cognitive deficits, while damage to local nodes does not decrease modularity and causes more isolated deficits, such as hemiplegia, or aphasia26,27. While connector hubs are not likely critical for only one specific cognitive process, their functions and diverse connectivity are required to maintain a cognitively optimal modular structure across cognitive processes. Thus, as we observed here, individual differences in the diversity of connector hubs' connectivity is predictive of cognitive performance across a range of very different tasks. Although diversely connected connector hubs are critical for successful performance in many different tasks, any given task nevertheless recruits very different cognitive and neural processes; each task likely engages connectivity patterns that are specifically optimal to that task. Future analyses should seek to understand both the general optimal connectivity patterns of connector hubs found here and the connectivity patterns that are optimal to a single task, including if and how these connectivity patterns interact. Methods Data and Preprocessing We used fMRI data from the Human Connectome Project48 S500 release. For the task-based fMRI data, Analysis of Functional NeuroImages (AFNI) was used to preprocess the images52. The AFNI command 3dTproject was used, passing the mean signal from the cerebral spinal fluid mask, the white matter mask, the whole brain signal, and the motion parameters to the "-ort" options, which removes these signals via linear regression. Within AFNI, the "-automask" option was used to generate the masks. The "-passband 0.009 0.08" option, which removes frequencies outside of 0.009 and 0.08, was used. Finally, the "-blur 6" option, which smooths the images (inside the mask only) with a 6mm FWHM filter after the time series filtering. Given the short length of the Emotion task (176 frames; Resting-State:1200, Social: 274, Relational:232, Motor:284, Language:316, Working Memory:405, Gambling:253) it was not included in our analyses. For the fMRI data collected at rest, we used the images that were previously preprocessed by the Human Connectome Project with ICA- FIX. We also used the AFNI command 3dBandpass to further preprocess these images. We used it to remove the mean whole brain signal and frequencies outside 0.009 and 0.08 (explicitly, "-ort whole_brain_signal.1D -band 0.009 0.08 -automask"). We did not regress out stimulus or task effects from the time series of each node, because how nodes' low frequency oscillations respond to stimulus or task effects is meaningful. Moreover, other investigators have noted that task effect regression has minimal effects53. As subject head motion during fMRI can impact functional connectivity estimates and has been shown to bias brain-task performance relationships54, performance prediction analyses were executed with scrubbing (removing frames with high motion) executed on frames with frame-wise displacement greater than 0.2 millimeters, including the frame before and after the movement. Frame-wise displacement measures movement of the head from one volume to the next, and was computed as the sum of the absolute values of the differentiated rigid body realignment estimates (translation and rotation in x, y, and z directions) at every time point with rotation values evaluated with a radius of 50 mm54. Frames were removed after all preprocessing was executed. Subjects with more than 75 percent of frames removed were not analyzed. Moreover, we executed all analyses after regressing out mean frame-wise displacement from the task performance values (Supplementary Figure 2). Graph Theory Analyses The Power atlas49 was used to define the 264 nodes in our graph because it was the only atlas that met all of the following requirements: (1) Given that the homogeneity of nodes in this atlas is high and they do not share physical boundaries, it will not overestimate the local connectivity of regions, (2) it is the only atlas that is defined based both on functional connectivity and studies of task activations making it optimal for our current analyses, (3) it accurately divides nodes into communities observed with other approaches (e.g., at the voxel level), and this division has been used in many studies33,36,49,55. A canonical division of nodes into communities aides in the interpretation and generalizability of our results. It can be found at: http://www.nil.wustl.edu/labs/petersen/Resources_files/Consensus264.xls. Moreover, we used this division to calculate within and between community edge weight changes across subjects. (4) It has anatomical coverage of cortical, subcortical, and cerebellar regions. All graph theory analyses were executed with our own custom python code (www.github.com/mb3152/brain_graphs) that uses the iGraph library. All analysis code is also publicly available (github.com/mb3152/hcp_performance/). For each task (both LR and RL encoding directions were used) and for each subject, the mean signal from 264 regions in the Power atlas was computed. The Pearson r between all pairs of signals was computed to form a 264 by 264 matrix, which was then Fisher z transformed. We chose Pearson r values to represent functional connectivity (i.e., edges) between nodes, for its simplicity in interpretation and ubiquity in human network neuroscience56. However, more complex statistical measures could be employed, including measures that attempt to estimate the directionality of each edge. The LR and RL matrices were then averaged. The mean matrix was then thresholded, retaining edge weights, at a range of costs (0.05 to 0.15 at 0.01 intervals), a common range and interval in graph theory analyses1,27,33,49. The maximum spanning tree was calculated to ensure all nodes had at least one edge. No negative correlations were included in our analyses. The matrix was then normalized to sum to a common value across subjects, and was used to represent the edges in the graph. Thus, all graphs had the same number of edges and sum of edge weights. For each cost, the InfoMap algorithm57 was run. While this method has been shown to be highly accurate on benchmark networks with known community structures, it is still a heuristic, as community detection is NP-hard58. While InfoMap does not explicitly maximize Q, it has been shown to estimate community structure accurately in several test cases59, rendering the Q value, the participation coefficients, and within community strengths computed based on the community structure accurate and valid. Moreover, in biological networks, InfoMap achieves Q values that are similar to algorithms that maximize Q 60; in the current resting-state data, InfoMap Q values and Fast-Greedy Q61 values were correlated at Pearson r=0.87 (dof=474, p<0.001, 95% CI: 0.84, 0.89); InfoMap Q values were found to be higher than Fast-Greedy Q values with a student's independent t-test (t(dof:952):16.027, p:<0.001, Cohen's d:0.775, 95% CI:0.024,0.03). InfoMap Q values and Louvain Q 62 values were correlated at Pearson r=0.98 (dof=474, p<0.001, 95% CI: (0.97, 0.98)); Louvain Q values were higher than InfoMap Q values. When comparing InfoMap Q values to the distribution that includes both Louvain and Fast-Greedy, two algorithms that explicitly maximize Q, InfoMap Q values were shown to be significantly higher with a student's independent t-test (t(dof:1429):5.304, p:<0.001, Cohen's d:0.222, 95% CI:0.005,0.011). Regardless, we found that it detects a community structure with Q values highly similar to other methods (Supplementary 𝛿8𝑐",𝑐#<. And modularity (Q) can be written as: Figure 15). Moreover, previous work has demonstrated the stability of community detection and the participation coefficient across community detection methods5. The participation coefficients, within community strengths, and Q were calculated at each cost. Q is written analytically as follows. Consider a weighted and undirected graph with n nodes and m edges represented by an adjacency matrix 𝐀 with elements 𝐀"#=edge weight between i and j. Thus, the strength of a node is given by 𝑘"=3𝐀"# # 𝑄= 12𝑚38𝐀"#−𝛾𝑝"#< "=# Here, 𝑝"# is the probability that nodes i and j are connected in a random null network 𝑃"#=𝑘"𝑘#2𝑚, γ is the resolution parameter, and 𝑐" is the community to which node 𝑖 belongs to and 𝛿(𝛼,𝛽)=1 if 𝛼=𝛽 and 𝛿(𝛼,𝛽)=0 if 𝛼≠𝛽. 𝑃𝐶"= 1− 3J𝐾"L𝐾"MN OP LQR Given a particular community assignment, the participation coefficient of each node can be calculated. The participation coefficient (PC) of node i is defined as: where 𝐾" is the sum of i 's edge weights, 𝐾"Lis the sum of i 's edge weights to community s, and NM is the total number of communities. Thus, the participation coefficient is a measure of how evenly distributed a node's edges are across communities. A node's participation coefficient is maximal if it has an equal sum of edge weights to each community in the network. A node's participation coefficient is 0 if all of its edges are to a single community. Finally, we calculate the within community strength value for each node as follows: 𝑧"=𝑘"−𝑘TLU 𝜎WXU Where 𝑘" is the number of links of node 𝑖 to other nodes in its community 𝑠", 𝑘TLU is the average of 𝑘 over all the nodes in 𝑠", and 𝜎WXU is the standard deviation of 𝑘 in 𝑠". Thus, The within community strength measures how well-connected node 𝑖 is to other nodes in the community relative to other nodes in the community. Each subject's participation coefficient, within community strength, and Q were the mean of those values across the range of costs. All analyses were executed and all prediction models were fit separately for each task. Tasks The following descriptions for each task have been adapted for brevity from the Human Connectome Project Manual63. Working Memory. The category specific representation task and the working memory task are combined into a single task paradigm. Participants were presented with blocks of trials that consisted of pictures of places, tools, faces and body parts (non-mutilated parts of bodies with no "nudity"). Within each run, the 4 different stimulus types were presented in separate blocks. Also, within each run, 1⁄2 of the blocks use a 2-back working memory task and 1⁄2 use a 0-back working memory task (as a working memory comparison). A 2.5 second cue indicates the task type (and target for 0-back) at the start of the block. Each of the two runs contains 8 task blocks (10 trials of 2.5 seconds each, for 25 seconds) and 4 fixation blocks (15 seconds). On each trial, the stimulus is presented for 2 seconds, followed by a 500 ms inter-task interval (ITI). Gambling. Participants play a card guessing game where they are asked to guess the number on a mystery card (represented by a "?") in order to win or lose money. Participants are told that potential card numbers range from 1-9 and to indicate if they think the mystery card number is more or less than 5 by pressing one of two buttons on the response box. Feedback is the number on the card (generated by the program as a function of whether the trial was a reward, loss or neutral trial) and either: 1) a green up arrow with "$1" for reward trials, 2) a red down arrow next to -$0.50 for loss trials; or 3) the number 5 and a gray double headed arrow for neutral trials. The "?" is presented for up to 1500 ms (if the participant responds before 1500 ms, a fixation cross is displayed for the remaining time), following by feedback for 1000 ms. There is a 1000 ms ITI with a "+" presented on the screen. The task is presented in blocks of 8 trials that are either mostly reward (6 reward trials pseudo randomly interleaved with either 1 neutral and 1 loss trial, 2 neutral trials, or 2 loss trials) or mostly loss (6 loss trials pseudo- randomly interleaved with either 1 neutral and 1 reward trial, 2 neutral trials, or 2 reward trials). In each of the two runs, there are 2 mostly reward and 2 mostly loss blocks, interleaved with 4 fixation blocks (15 seconds each). Motor. Participants are presented with visual cues that ask them to either tap their left or right fingers, or squeeze their left or right toes, or move their tongue to map motor areas. Each block of a movement type lasted 12 seconds (10 movements), and is preceded by a 3 second cue. In each of the two runs, there are 13 blocks, with 2 of tongue movements, 4 of hand movements (2 right and 2 left), and 4 of foot movements (2 right and 2 left). In addition, there are 3 15-second fixation blocks per run. Language & Math. The task consists of two runs that each interleave 4 blocks of a story task and 4 blocks of a math task. The lengths of the blocks vary (average of approximately 30 seconds), but the task was designed so that the math task blocks match the length of the story task blocks, with some additional math trials at the end of the task to complete the 3.8 minute run as needed. The story blocks present participants with brief auditory stories (5-9 sentences) adapted from Aesop's fables, followed by a 2-alternative forced- choice question that asks participants about the topic of the story. For example: "after a story about an eagle that saves a man who had done him a favor, participants were asked, "Was that about revenge or reciprocity?" The math task also presents trials auditorily and requires subjects to complete addition and subtraction problems. The trials present subjects with a series of arithmetic operations (e.g., "fourteen plus twelve"), followed by "equals" and then two choices (e.g., "twenty- nine or twenty- six"). Participants push a button to select either the first or the second answer. The tasks are adaptive to try to maintain a similar level of difficulty across participants. Social (Theory of Mind). Participants were presented with short video clips (20 seconds) of objects (squares, circles, triangles) that either interacted in some way, or moved randomly on the screen. After each video clip, participants judge whether the objects had a mental interaction (an interaction that appears as if the shapes are taking into account each other's feelings and thoughts), Not Sure, or No interaction (i.e., there is no obvious interaction between the shapes and the movement appears random). Each of the two task runs has 5 video blocks (2 Mental and 3 Random in one run, 3 Mental and 2 Random in the other run) and 5 fixation blocks (15 seconds each). Relational. The stimuli are 6 different shapes filled with 1 of 6 different textures. In the relational processing condition, participants are presented with 2 pairs of objects, with one pair at the top of the screen and the other pair at the bottom of the screen. They are told that they should first decide what dimension differs across the top pair of objects (differed in shape or differed in texture) and then they should decide whether the bottom pair of objects also differ along that same dimension (e.g., if the top pair differs in shape, does the bottom pair also differ in shape). In the control matching condition, participants are shown two objects at the top of the screen and one object at the bottom of the screen, and a word in the middle of the screen (either "shape" or "texture"). They are told to decide whether the bottom object matches either of the top two objects on that dimension (e.g., if the word is "shape", is the bottom object the same shape as either of the top two objects. For both conditions, the subject responds yes or no using one button or another. For the relational condition, the stimuli are presented for 3500 ms, with a 500 ms ITI, and there are four trials per block. In the matching condition, stimuli are presented for 2800 ms, with a 400 ms ITI, and there are 5 trials per block. Each type of block (relational or matching) lasts a total of 18 seconds. In each of the two runs of this task, there are 3 relational blocks, 3 matching blocks and 3 16-second fixation blocks. Performance measures. All performance measures were chosen a priori. In the working memory task, we used the mean accuracy across all n-back conditions (face, body, place, tool). In the relational task, we used mean accuracy across both the matching and the relational conditions. For the language task, we took the maximum difficulty level that the subject achieved across both the math and language conditions. We did not use accuracy, because the task varies in difficulty based on how well the subject is doing, making accuracy an inaccurate measure of performance for these tasks. For the social task, given that almost all subjects correctly identified the social interactions as social interactions, we used the percentage of correctly identified random interactions. Deep neural network model. A deep neural network is a supervised learning algorithm that can learn a non-linear function for regression or classification. Unlike logistic regression, there are one or more non-linear layers, called hidden layers, between the input and the output layer. Thus, the model is trained to relate a set of input features to outputs by learning weights between neurons across adjacent layers (Supplementary Figure 1). Our implementation uses the sklearn python library. Explicitly, a prediction for subject z is calculated as: model = sklearn.neural_network.MLPRegressor(hidden_layer_sizes=(8,12,8,12)) model.fit(x[subjects!=z], y[subjects!=z]) prediction = model.predict(x([z]) where x is the set of features across subjects and y is the task performance across subjects. Analytic quality of diversity and locality facilitated modularity coefficients To further understand diversity and locality facilitated modularity coefficients, we performed an iterative split-half analysis. Specifically, we estimated the mean within community strength or participation coefficient of each node in one half of subjects, and each node's locality and diversity facilitated coefficient in the other half, testing 10,000 random splits of subjects. All relationships were reliably observed in every cognitive state (Supplementary Figure 7). Next, we sought to determine if this relationship was a necessary feature of the underlying mathematics, or whether it was a phenomenon specific to the neurophysiology of brain networks. To address this question, we tested four null model networks and observed that none of them exhibited a significant relationship between mean participation coefficient and diversity facilitated modularity coefficient (Supplementary Figure 8). As a third check, we assessed whether the number of communities identified in the network was inadvertently biasing our results. We observed that the number of communities in each network was negatively correlated with the modularity value Q (Supplementary Figure 9). After regressing out the number of communities in each network from the modularity value, we observed that our findings remained qualitatively unchanged (Supplementary Figure 10). Finally, we tested whether the relationships between variables of interest were linear (and therefore appropriate to examine with Pearson r correlation coefficients), or nonlinear. To address this question, we analyzed individual 1st, 2nd, and 3rd order curve fits of the relationship between participation coefficients and modularity values. We observed that many relationships were well-captured by a first order fit, with the connector hub's maximal participation coefficients corresponding to maximal Q indices, with only a few showing a more nonlinear relationship (Supplementary Figure 11). Alternative analysis of connector hubs' tuning the connectivity of their neighbors We executed an alternative analysis to test if connector hubs tune the connectivity of their neighbors to be more modular. For each node i we calculated a matrix, where the j-kth entry is the Pearson r correlation coefficient that captures how well the participation coefficients of node i correlates with the edge weights between nodes j and k in the network across subjects. These Pearson r values allowed us to test whether a node's participation coefficients correlate positively with its neighbors' increased connectivity to its own community and decreased connectivity to other communities. We subtracted the sum of r values in the matrix corresponding to node i's participation coefficients and node j's between community edge weights from the sum of r values in the matrix corresponding to node i's participation coefficients and node j's within community edge weights. Thus, this value measures how well the participation coefficients of node i are correlated with the increased modular (within community) connectivity of node j. We used the partition of nodes into communities that was created along with the nodes themselves (Figure 6d)49. Edges between node i and node j were ignored in this calculation, as the participation coefficients of node i is likely highly correlated with the edge weights between node i and node j, and we were only interested in how the participation coefficient of node i modulates node j's connectivity to the rest of the network, not node j's connectivity to node i. Edges that were not positive on average across subjects were not included in this analysis, as the interpretation of negative edges in fMRI-based networks is not obvious (results were similar only including the top 25 percent of edges (Supplementary Figure 13). Correlations between the edge strength between nodes i and j and the amount of modulation of j's modularity by i were calculated such that the set of nodes i were either connector hubs or non-connector hubs. A positive correlation means that a node is biased to modulate the connectivity of its neighbors versus its non-neighbors to be more modular. In all cognitive states, these correlations were only positive and significant (Pearson's r>0.17, p<0.001, Bonferroni corrected (n tests = 7)) for connector hubs (Supplementary Figure 13) suggesting that connector hubs tune the connectivity of their neighbors to be more modular. To test if this relationship existed for local hubs, we calculated a similar matrix, where, for each node i, the j-k th entry is the Pearson r value that captures how well the within community strengths of node i correlate with the edge weight between nodes j and k in the network across subjects. Correlations between the edge strength between nodes i and j and the amount of modulation of j's modularity by the within community strength of i were calculated such that nodes i were either local hubs or non-local hubs. None of these correlations were robust (-0.1>r<0.1). These analyses add to our conclusion, demonstrated in the Results, that connector hubs facilitate higher modularity by tuning the connectivity of their neighbors to be more modular. Statistical Methods The number of subjects was determined by the number of subjects released by the Human Connectome Project at the start of the analyses. As this dataset represented the largest dataset of its kind at that time and the number of subjects is greater than many similar analyses47, no power analysis was computed. Total N=Working Memory: 475, Gambling: 473, Relational:458, Motor:475, Language & Math:472, Social:474, Resting State: 476. However, as we only analyzed subjects with both Resting-State and the task scans, N=Working Memory:473, Relational:457, Language & Math:471, Social:473. This results in a unique N=476 across tasks, in that 476 different subjects had a resting state scan and at least one task scan. As scrubbing (which removes frames with large head motion) can cause too many frames to be removed from the time series, subjects with less than 75 percent of remaining frames were not included in the analyses that implemented scrubbing; thus, for analyses using scrubbed data, N=Working Memory:351, Relational:335, Language & Math:348, Social:358. All confidence intervals (CI) are reported with alpha=0.05. For Pearson r correlation coefficients CIs, the interval of r values is given by Fisher transforming r to z, computing the interval, and then Fisher reverse transforming the z intervals back to r intervals. For t-tests, the confidence interval represents the largest and smallest differences in means across the two distributions. For all t-tests, distributions were confirmed as normal (p<0.001) or exhibiting no significant evidence as not normal (k2>0.0) using D'Agostino and Pearson's omnibus test k2. All p values are two sided tests. All p values that are part of a family of tests are Bonferroni corrected for multiple comparisons. For example, we test if two tasks' hub and network structures are similarly optimal for the same subject measures, testing across a large number of subject measures. In this case, we applied a Bonferroni correction to the p-values to determine whether the effect remained true for particular subject measures. Here, the number of tests is equal to the number of subject measures, 47. Individual subject networks were built independently for each task and task performance is different for each task. Thus, these tests are not strictly in the same family. However, to be conservative, we still Bonferroni corrected these p-values. In these cases, the family size is either 4 or 7, depending on the number of tasks analyzed. Unless otherwise stated, all p values are Bonferroni corrected. Many statistical tests are calculated here without reported p values. For example, Pearson r values are used to calculate functional connectivity. Here, only the r values are of interest -- more precisely, individual differences in the r values across subjects, and how these differences relate to individual differences in cognition. This treatment of multiple comparisons in the context of functional connectivity and individual differences in cognition is common and recommended47,64. We extend this notion to other analyses here as well. For example, we use the Pearson correlation coefficient r to compare how well different nodes' participation coefficients across subjects explain variance in network modularity or task performance (the diversity facilitated modularity and performance coefficients). In these cases, we relate these r-values to other measures, and are only concerned with how these r-values explain another distributions' variance (here, we find a positive correlation between these r-values and a node's mean participation coefficient across subjects). We are not concerned with the statistical significance any particular r-value as estimated by the p-value. We care about the distribution of r-values, not the distribution of p-values, and we do not make any claims about any single r-value. Thus, the p-values are neither reported nor corrected for multiple comparisons. This is precisely how functional connectivity is treated statistically. Acknowledgements This work was supported by NIH Grant NS79698 and the National Science Foundation Graduate Research Fellowship Program under Grant no. DGE 1106400 to MAB and MD. MAB would also like to acknowledge NIH T32 Ruth L. Kirschstein Institutional National Research Service Award (5T32MH106442-02). BTTY was also supported Singapore MOE Tier 2 (MOE2014-T2-2-016), NUS Strategic Research (DPRT/944/09/14), NUS SOM Aspiration Fund (R185000271720), Singapore NMRC (CBRG/0088/2015), NUS YIA and the Singapore National Research Foundation (NRF) Fellowship (Class of 2017). DSB would also like to acknowledge support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Research Laboratory and the Army Research Office through contract numbers W911NF-10-2-0022 and W911NF-14-1-0679, the National Institute of Health (2-R01- DC-009209-11,1R01HD086888-01, R01-MH107235, R01-MH107703, and R21-M MH- 106799), the Office of Naval Research, and the National Science Foundation (BCS- 1441502, PHY-1554488, and BCS-1631550). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. Data availability Data were provided by the Human Connectome Project, WU-Minn Consortium (Principal Investigators: David Van Essen and Kamil Ugurbil; 1U54MH091657) funded by the 16 NIH Institutes and Centers that support the NIH Blueprint for Neuroscience Research and by the McDonnell Center for Systems Neuroscience at Washington University. The content is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies. All analyses were executed in accordance with the authors' institutions' relevant ethical regulations as well as the WU-Minn HCP Consortium Open Access Data Use Terms. Informed consent was obtained from all participants. Competing Interests The authors declare no competing interests. Code availability All graph theory analyses were executed with our own custom python code (www.github.com/mb3152/brain_graphs) that uses the iGraph library. All analysis code is also publicly available (github.com/mb3152/hcp_performance/). Author Contributions M.A.B. conceived the analyses. M.A.B., B.T.T.Y., D.S.B., and M.D. collaboratively designed the analyses. M.A.B. executed the analyses. M.A.B., B.T.T.Y., D.S.B., and M.D. collaboratively wrote the paper. Supplementary Figure 1. Workflow of the deep neural network features and model construction. a, For each node, we measured the diversity facilitated performance coefficient. For example, the Pearson r is calculated between node 9's participation coefficients across subjects (black outline) and task performance across subjects (black outline), resulting in a diversity facilitated performance coefficient of 0.23 (black outline) for that node. b, The optimality of a subject's participation coefficients for task performance is measured by, for example, calculating the Pearson r between subject 9's participation coefficients (black outline) and the diversity facilitated performance coefficients (black outline), resulting in a diversity feature for the model of 0.43 (black outline). The same procedures in a and b are also executed using within community strengths and edge weights instead of participation coefficients. Modularity (Q) values are also used as features. These four features are derived separately from resting-state and the task network for which task performance is being predicted, resulting in eight features. All calculations in panels a and b are calculated without data from the subject for which the prediction is being made. These eight features are then used in a deep neural network model (c) to learn the relationship between the features and task performance (again, without data from the subject for which the prediction is being made) by adjusting the weights between nodes in adjacent layers. The features are then calculated for the left-out subject (e.g., Pearson r between the left-out subject's participation coefficients and the previously calculated diversity facilitated performance coefficients that did not include data from the left-out subject) and are used in the deep neural network model to generate a prediction for the left-out subject. Supplementary Figure 2 Hub diversity and locality, modularity, and network connectivity predicts task performance. a, predictive model of task performance (see Methods for performance measures), as shown in Figure 2a. b, Predictive model of task performance without using network connectivity (i.e., only hub diversity and locality and modularity). N=Working Memory: 473, Relational: 457,Language & Math: 471, Social: 473). Each dot represents a subject's prediction. Shaded areas represent 95 percent confidence intervals. c,d, As in a,b, except the frames with high motion were removed from the time series before the network of that individual was constructed (i.e., scrubbing), N=Working Memory: 351, Relational: 335, Language & Math: 348, Social: 358, as subjects with less than 75 percent of frames after scrubbing were not analyzed. e,f, As in a,b, except mean frame wise displacement was regressed out from the task performance values, N=Working Memory: 473, Relational: 457,Language & Math: 471, Social: 473). All p values are Bonferroni corrected (n tests=4). Supplementary Figure 3. Feature correspondence between tasks and subject measures. To test if certain hub and network structures that are optimal to each task (Language & Math, Relational, Social, or Working Memory) are also optimal for other subject measures, we measured the feature correspondence between subjects' features in the task model with the subjects' features in the same model that was fit to a given subject measure (e.g., Delayed Discounting) instead of performance in the task. The feature correspondence is shown for each task; high feature correspondence means a similar hub and network structure is optimal for the task and subject measure. Colors represent z-scored (within column) feature correspondence values. Results significant at p<1-e3 uncorrected, p<0.05 Bonferroni corrected (N tests= 47) are marked with an asterisk. N subjects = Working Memory: 473, Relational: 457,Language & Math: 471, Social: 473). Supplementary Figure 4 Prediction of subject measures by the model of hub diversity and locality, modularity, and network connectivity. As in Figure 1, we predicted each subject measure using a model of hub diversity and locality, modularity, and network connectivity. This was executed for networks constructed based on the data from each task (columns). For each task and subject measure, the Pearson r between the real subject measure values and the predicted subject measure values are shown. Colors represent the z-scored (within each task) Pearson r. Predictions that are significant (p<1-e3 uncorrected, p<0.05 after Bonferroni correction, N tests = 47) are marked with an asterisk. N= Working Memory: 473, Relational: 457, Language & Math: 471, Social: 473). Supplementary Figure 5 Analyzing individual differences of hubs' diversity and locality and modularity. When connector hubs are connected to many communities (as indicated by high participation coefficients), they can integrate information and tune connectivity optimally, allowing other regions to perform more modular local processing. Thus, across subjects, we predicted that the increased participation coefficients (i.e., diversity) of connector hubs would be correlated with preservations or increases in the modularity (Q) of the network. This prediction is illustrated above (a). Each graph represents the network structure of an individual subject. A single connector hub, shown in red, is identified based on its high mean participation coefficient across individuals; subject-level participation coefficient values are shown inside the red nodes. Local neighborhoods or communities are shown in green, blue, and pink, and subject-level values of modularity(Q) are shown above each graph. Across subjects, we predict that higher participation coefficients of connector hubs will occur in more modular networks in which the neighbors of connector hubs (pink, dashed-outline) display more local connectivity. b, To test this prediction, for each node we measured the diversity facilitated modularity coefficient; for example, the Pearson r is calculated between node 9's participation coefficients across subjects (black outline) and modularity (Q) across subjects (black outline), resulting in a diversity facilitated modularity coefficient of 0.27 (black outline). Supplementary Figure 6 Connector and local hubs facilitate network modularity and task performance. a, correlations between a node's mean participation coefficient (across subjects) and the node's diversity facilitated modularity coefficient (the Pearson r correlation coefficient between that node's participation coefficients and modularity values (Q) across subjects). Correlations were calculated for each cognitive state. Each dot represents a node in the brain's functional network. Shaded areas represent 95 percent confidence intervals. In all cognitive states, there was a significant positive correlation between a node's mean participation coefficient and the node's diversity facilitated modularity coefficient. This demonstrates that connector hubs facilitate increased modularity. b, correlations between a node's mean within community strength (across subjects) and the node's locality facilitated modularity coefficient (Pearson r correlation coefficient between that node's within community strengths and modularity values (Q) across subjects). In all cognitive states, there was a significant positive correlation between a node's within community strength and the node's locality facilitated modularity coefficient. This demonstrates that local hubs facilitate increased modularity. c, correlations between a node's mean participation coefficient (across subjects) and the node's diversity facilitated performance coefficient (the Pearson r correlation coefficient between that node's participation coefficients and task performance across subjects). Correlations were calculated for each cognitive state. Each dot represents a node in the brain's functional network. Shaded areas represent 95 percent confidence intervals. In all cognitive states, there was a significant positive correlation between a node's mean participation coefficient and the node's diversity-facilitated performance coefficient. This demonstrates that connector hubs facilitate increased task performance. d, correlations between a node's mean within community strength (across subjects) and the node's locality facilitated performance coefficient (Pearson r correlation coefficient between that node's within community strengths and task performance across subjects). In all cognitive states (except Relational, Bonferroni (n tests=4) p=0.081, uncorrected p=0.02), there was a significant positive correlation between a node's within community strength and the node's locality facilitated performance coefficient. Bonferroni corrected p values are shown (n tests=7(a,b), and 4(c,d). N=264, the number of nodes in the graph. This demonstrates that local hubs facilitate increased task performance. Supplementary Figure 7 Cross validation of diversity and locality facilitated modularity coefficients. Connector hubs and local hubs are defined by a higher participation coefficient and within community strength, respectively, across subjects on average. To avoid potential dependencies, we estimated the mean within community strength or participation coefficient of each node in one half of subjects, and the correlation between a nodes' within community strength or participation coefficients and Q indices -- the node's locality and diversity facilitated modularity coefficients -- in the other half, testing 10,000 split of subjects. On average, correlations were still significant in every state. a, The distribution of Pearson r values between a node's mean participation coefficient (defined in half the subjects) and the diversity facilitated modularity coefficient (defined in the other half of the subjects). b, The distribution of Pearson r values between a node's mean within community strength (defined in half the subjects) and the locality facilitated modularity coefficient (defined in the other half of the subjects). N=264, the number of nodes in the graph, for each r value calculation, (dof=262). The mean and quartiles are marked in each violin. Supplementary Figure 8 Null Models. We tested four null models of diversity facilitated modularity coefficients. (1) Random Edges, Real Community utilizes the true partition of nodes into communities that was uncovered by the application of community detection to each subject's intact resting-state graph (thresholded at a weighted graph density of 0.10, chosen based on this cost being the median cost from our original analyses), but we randomly permuted the edges uniformly after the partition was identified. (2) Random Community, Real Edges utilizes the partition of nodes into communities that was uncovered by the application of community detection to each subject's intact resting-state graph (thresholded at a weighted graph density of 0.10), but we then randomly permuted the assignment of nodes to communities (this retains the same number of communities and sizes of communities as the true partition). The edges remain in their true locations. (3) Random Edges, Clustered permutes the edges in each subject's resting-state graph uniformly at random. We then applied community detection to this permuted graph to identify a partition of nodes into communities, and then calculated the participation coefficient of each node based on that partition. We used a cost of 0.05, as denser random graphs result in just one community, which results in trivial participation coefficients of 0. For each instantiation (n=100) of each of these three models, we generated a graph for each subject using the subject's original graph and that particular null model. Finally, (4) we generated Wattz-Strogatz small world graphs. Each graph had 264 nodes with each node initially connected to its 7 neighbors in the lattice. We set the rewiring probability to 0.25. This results in Q values of roughly 0.40 and a binary density of roughly 0.05. 100 graphs were generated for each instantiation(n=100) of the model. For each null model, we then calculated the diversity facilitated modularity coefficient. Violin plots are shown for the distribution of the correlation between each node's mean participation coefficient and each node's diversity facilitated modularity coefficient across 100 instantiations of each model. For comparison, the distribution from the original analysis across tasks (Real Edges, Real Community) is also shown. N=264, the number of nodes in the graph, dof=262. Supplementary Figure 9 Participation coefficients, the number of communities in the network, and modularity (Q). a, Negative correlation between the number of communities in subjects' graphs (averaged across graph densities in the range 0.05- 0.15) and the Q value of the graphs (also averaged across graph densities in the range 0.05-0.15). b, We performed a mediation analysis between each node's participation coefficients, the number of communities in the graphs, and the Q values of the graphs, with the number of communities being the mediator. Mediation values are plotted for each node on the x-axis. The y-axis is the Pearson r between each node's participation coefficients and Q indices across subjects' graphs -- the diversity facilitated modularity coefficient. Shaded areas represent a 95 percent confidence interval. N=264, the number of nodes in the graph, dof=262. Supplementary Figure 10 Diversity and locality facilitated modularity coefficients controlling for the number of communities. a, Correlations between a node's mean participation coefficient (across subjects) and the node's diversity facilitated modularity coefficient (the Pearson r correlation coefficient between that node's participation coefficients and modularity values (Q) across subjects after regressing out the number of communities in the network from Q). Correlations were calculated for each cognitive state. Each dot represents a node in the brain's functional network. Shaded areas represent 95 percent confidence intervals. In all cognitive states, there was a significant positive correlation between a node's mean participation coefficient and the node's diversity-facilitated modularity coefficient. b, Correlations between a node's mean within community strength (across subjects) and the node's locality facilitated modularity coefficient (Pearson r correlation coefficient between that node's within community strength and modularity values (Q) across subjects after regressing out the number of communities in the network from Q. In all cognitive states (except Resting state (uncorrected p=0.0132, Bonferroni corrected p=0.093), there was a significant positive correlation between a node's within community strength and the node's locality facilitated modularity coefficient. All p values are Bonferroni corrected (n tests=7). N=264, the number of nodes in the graph. N subjects=Working Memory:475, Gambling:473, Relational: 458, Motor:475, Language:472, Social: 474, Rest: 476. Supplementary Figure 11 Relationships between individual connector hubs' participation coefficients and Q. To test if the relationship between a connector hub's participation coefficients and Q -- the connector hubs' diversity facilitated modularity coefficient -- was linear, we fit three regression models to individual connector hubs' participation coefficients and Q values across subjects. In each plot, each line is the relationship between a single connector hub's participation coefficients and Q values across subjects. Only nodes with positive Pearson r values are shown. a, Locally weighted scatter-plot smoother fit. b, 2nd order fit. c, 3rd order fit. The relation between many connector hubs' participation coefficients and Q indices across subjects is well captured by a first order fit, with the hub's maximal participation coefficients (0.7; the mathematically upper limit is 0.99T) corresponding to maximal Q. However, for example, some connector hubs' participation coefficients correspond to the maximal Q at a participation coefficient of 0.4-0.5, and then Q decreases at higher participation coefficients of that connector hub. Shaded areas represent 95 percent confidence intervals. N=476. Supplementary Figure 12 Connector hubs' mediation of neighboring nodes' edges. To investigate if the relationship between a connector hub's participation coefficient and Q is mediated primarily by that connector hub's neighbors' edge pattern increasing Q (per our prediction), we executed, for each connector hub node (i), a t-test between the absolute mediation values of node (i)'s neighbors' edges versus the absolute mediation values of node (i)'s non-neighbors' edges (neighbors were defined based on edges present between the two nodes in a graph at a density of 0.15, which was used because it is the densest density we utilized in our analyses; neighbors' and non-neighbors' edges connecting to the node (i) were ignored). Mediation values were based on the edge mediating between node (i)'s participation coefficients and modularity values (Q); the distributions of mediation values was tested for normality using D'Agostino and Pearson's omnibus test k2. All distributions were confirmed as normal (k2>100.0, p<0.001 for all tasks). The distribution of t-values for connector hubs is shown for each task. The mean quartiles are marked. In general, connector hubs showed higher mediation values with edges of its neighbors compared to the edges of its non-neighbors (i.e., t>0). Moreover, these t-values were significantly higher than the same t-values calculated with local hubs (t(dof:1358): 3.892, p:0.0001, Cohen's d:0.219, CI:1.887,6.62). Supplementary Figure 13 Connector hubs' relationships with individual edges' weights. a,b, For each task, for each pair of nodes i and j, we calculated the correlation between (x) how strongly node i increased the within community edge strength of node j (the sum of Pearson r values between the participation coefficients of node i and the within community edges of node j minus the sum of Pearson r values between the participation coefficients of node i and the between community edges of node j) with (y) the average connectivity edge weight between node i and node j. a Includes all positive edges, while b includes the strongest 25 percent of edges. Bonferroni corrected p values are shown in all plots (n tests = 7). These results show that connector hubs are predominately tuning the connectivity of their neighbors. Shaded areas represent 95 percent confidence intervals. N=dof+2 for each panel. Supplementary Figure 14 Connector hubs and local hubs are interrelated. For each task, across subjects, the correlation between the subject's connector hubs' mean participation coefficient and the subject's local hubs' mean within community strengths. In every task, if a subject's connector hubs were diverse, local hubs were local. Here, connector and local hubs were defined based on nodes for which their diversity and locality (respectively) facilitated modularity coefficients were positive. For this calculation, participation coefficients were z-scored within each subject, as within community strengths are z-scored within each subject. Bonferroni corrected p values are reported in all plots(n tests = 7). Shaded areas represent 95 percent confidence intervals. N=Working Memory:475, Gambling:473, Relational: 458, Motor:475, Language:472, Social: 474, Rest: 476. Supplementary Figure 15 Modularity quality indices from three community detection algorithms. We used Q indices repeatedly in our analyses. However, the community detection algorithm that we utilized, InfoMap, does not explicitly maximize Q. To see if this could potentially impact our analyses, we compared Q values from InfoMap to two popular algorithms, Fast-Greedy and Louvain, that explicitly maximize Q. The mean Q value, as in our analyses, was taken across costs of 0.05 to 0.15. a, Distribution, across subjects, of Q values of each algorithm. b, Correlation between Q values across algorithms. Shaded areas represent 95 percent confidence intervals. In all panels, N=476, dof=474. 2. 3. Guimerà, R., Guimera, R., Amaral, L. A. N. & Amaral, L. Functional cartography of 4. Meunier, D., Lambiotte, R. & Bullmore, E. T. Modular and hierarchically modular 5. 6. Bertolero, M. A., Yeo, B. T. T., & D'Esposito, M. The modular and integrative functional architecture of the human brain. Proc. Natl. Acad. Sci. U.S.A. 112, E6798 -- 807 (2015). Bertolero, M. A., Yeo, B. T. T., & D'Esposito, M. The diverse club. Nature Communications 8, 1277 (2017). complex metabolic networks. Nature 433, 895 -- 900 (2005). organization of brain networks. Front. Neurosci. 4, 200 (2010). Bertolero, M. A., Yeo, B. T. T., & D'Esposito, M. The diverse club. Nature Communications 8, 1 -- 10 (2017). Bassett, D. S. et al. Efficient Physical Embedding of Topologically Complex Information Processing Networks in Brains and Computer Circuits. PLoS Comput Biol 6, e1000748 (2010). Laughlin, S. B., de Ruyter van Steveninck, R. R. & Anderson, J. C. The metabolic cost of neural information. Nature Neuroscience 1, 36 -- 41 (1998). Lord, L.-D., Expert, P., Huckins, J. F. & Turkheimer, F. E. Cerebral Energy Metabolism and the Brain's Functional Network Architecture: An Integrative Review. Journal of Cerebral Blood Flow & Metabolism 33, 1347 -- 1354 (2013). Raichle, M. E. & Gusnard, D. A. Appraising the brain's energy budget. Proceedings of the National Academy of Sciences 99, 10237 -- 10239 (2002). Neuroscience 32, 356 -- 371 (2012). development. Proceedings of the National Academy of Sciences 111, 13010 -- 13015 (2014). 12. Krienen, F. M., Yeo, B. T. T., Ge, T., Buckner, R. L. & Sherwood, C. C. Transcriptional profiles of supragranular-enriched genes associate with corticocortical network architecture in the human brain. Proc. Natl. Acad. Sci. U.S.A. 113, E469 -- 78 (2016). 13. Hawrylycz, M. et al. Canonical genetic signatures of the adult human brain. Nature Neuroscience 18, 1832 -- 1844 (2015). 14. Wagner, G. P. & Zhang, J. The pleiotropic structure of the genotype-phenotype map: the evolvability of complex organisms. Nature Reviews Genetics 12, 204 -- 213 (2011). Biol. Sci. 280, 1 -- 9 (2013). modularity and network motifs. 102, 13773 -- 13778 (2005). doi:10.1007/978-1-4899-0718-9_31 15. Clune, J., Mouret, J.-B. & Lipson, H. The evolutionary origins of modularity. Proc. 16. Kashtan, N., Kashtan, N., Alon, U. & Alon, U. Spontaneous evolution of 17. Simon, H. A. in Facets of Systems Science 457 -- 476 (Springer US, 1991). 18. Fodor, J. & Fodor, J. The Modularity of Mind. (MIT Press, 1983). 19. Coltheart, M. Modularity and cognition. Trends in Cognitive Sciences 3, 115 -- 120 9. 10. Harris, J. J. & Attwell, D. The Energetics of CNS White Matter. Journal of 11. Kuzawa, C. W. et al. Metabolic costs and evolutionary implications of human brain 1. 7. 8. (1999). 20. Robinson, P. A., Henderson, J. A., Matar, E., Riley, P. & Gray, R. T. Dynamical 23. Stevens, A. A., Tappon, S. C., Garg, A. & Fair, D. A. Functional brain network 28. 29. Rubinov, M., Ypma, R. J. F., Watson, C. & Bullmore, E. T. Wiring cost and Reconnection and Stability Constraints on Cortical Network Architecture. Physical Review Letters 103, 108104 (2009). 21. Clune, J., Mouret, J.-B. & Lipson, H. The evolutionary origins of modularity. Proc. Biol. Sci. 280, 20122863 -- 20122863 (2013). 22. Tosh, C. R., Tosh, C. R., McNally, L. & McNally, L. The relative efficiency of modular and non-modular networks of different size. Proc. Biol. Sci. 282, 20142568 -- 20142568 (2015). modularity captures inter- and intra-individual variation in working memory capacity. PLOS ONE 7, e30468 (2012). 24. Arnemann, K. L. et al. Functional brain network modularity predicts response to cognitive training after brain injury. Neurology 84, 1568 -- 1574 (2015). 25. Modular Brain Network Organization Predicts Response to Cognitive Training in Older Adults. PLoS ONE (2016). 26. Warren, D. E. et al. Network measures predict neuropsychological outcome after brain injury. Proc. Natl. Acad. Sci. U.S.A. 111, 14247 -- 14252 (2014). 27. Gratton, C., Nomura, E. M., Pérez, F. & D'Esposito, M. Focal Brain Lesions to Critical Locations Cause Widespread Disruption of the Modular Organization of the Brain. Journal of Cognitive Neuroscience 24, 1275 -- 1285 (2012). van den Heuvel, M. P. & Sporns, O. Network hubs in the human brain. Trends in Cognitive Sciences 17, 683 -- 696 (2013). topological participation of the mouse brain connectome. Proc. Natl. Acad. Sci. U.S.A. 112, 10032 -- 10037 (2015). 30. Ferreira, F. R. M., Nogueira, M. I. & Defelipe, J. The influence of James and Darwin on Cajal and his research into the neuron theory and evolution of the nervous system. Front Neuroanat 8, 1 (2014). transportation network: Anomalous centrality, community structure, and cities' global roles. Proceedings of the National Academy of Sciences 102, 7794 -- 7799 (2005). 32. Guimerà, R., Sales-Pardo, M. & Amaral, L. A. N. Classes of complex networks defined by role-to-role connectivity profiles. Nature Physics 3, 63 -- 69 (2007). 33. Power, J. D., Schlaggar, B. L., Lessov-Schlaggar, C. N. & Petersen, S. E. Evidence for Hubs in Human Functional Brain Networks. 79, 798 -- 813 (2013). van den Heuvel, M. P. & Sporns, O. An Anatomical Substrate for Integration 34. among Functional Networks in Human Cortex. 33, 14489 -- 14500 (2013). 35. Scholtens, L. H., Schmidt, R., de Reus, M. A. & van den Heuvel, M. P. Linking Macroscale Graph Analytical Organization to Microscale Neuroarchitectonics in the Macaque Connectome. 34, 12192 -- 12205 (2014). 36. Cole, M. W. et al. Multi-task connectivity reveals flexible hubs for adaptive task control. Nature Neuroscience 16, 1348 -- 1355 (2013). 31. Guimera, R., Mossa, S., Turtschi, A. & Amaral, L. A. N. The worldwide air 43. Sala-Llonch, R. et al. Brain connectivity during resting state and subsequent 44. Sadaghiani, S., Poline, J.-B., Kleinschmidt, A. & D'Esposito, M. Ongoing 37. Yeo, B. T. T. et al. Functional Specialization and Flexibility in Human Association Cortex. Cereb Cortex 25, 3654 -- 3672 (2015). 38. Bassett, D. S., Yang, M., Wymbs, N. F. & Grafton, S. T. Learning-induced autonomy of sensorimotor systems. Nature Neuroscience 18, 744 -- 751 (2015). 39. Spadone, S. et al. Dynamic reorganization of human resting-state networks during visuospatial attention. Proc. Natl. Acad. Sci. U.S.A. 112, 8112 -- 8117 (2015). 40. Gratton, C., Laumann, T. O., Gordon, E. M., Adeyemo, B. & Petersen, S. E. Evidence for Two Independent Factors that Modify Brain Networks to Meet Task Goals. Cell Reports 17, 1276 -- 1288 (2016). 41. Yamashita, M., Kawato, M. & Imamizu, H. Predicting learning plateau of working memory from whole-brain intrinsic network connectivity patterns. Scientific Reports 5, 7622 (2015). 42. Baldassarre, A. et al. Individual variability in functional connectivity predicts performance of a perceptual task. Proc. Natl. Acad. Sci. U.S.A. 109, 3516 -- 3521 (2012). working memory task predicts behavioural performance. Cortex 48, 1187 -- 1196 (2012). dynamics in large-scale functional connectivity predict perception. Proc. Natl. Acad. Sci. U.S.A. 112, 8463 -- 8468 (2015). Network States during Cognitive Task Performance. Neuron 92, 544 -- 554 (2016). abstinence effects on craving and cognitive function. JAMA Psychiatry 71, 523 -- 530 (2014). 47. Finn, E. S. et al. Functional connectome fingerprinting: identifying individuals using patterns of brain connectivity. Nature Neuroscience 18, 1664 -- 1671 (2015). 48. Van Essen, D. C. et al. The WU-Minn Human Connectome Project: an overview. NeuroImage 80, 62 -- 79 (2013). 49. Power, J. D. et al. Functional Network Organization of the Human Brain. Neuron 72, 665 -- 678 (2011). 50. Smith, S. M. et al. A positive-negative mode of population covariation links brain connectivity, demographics and behavior. Nature Neuroscience 18, 1565 -- 1567 (2015). 51. Bullmore, E. & Sporns, O. The economy of brain network organization. Nat Rev Neurosci 74, 47 -- 14 (2012). 52. Cox, R. W. AFNI: software for analysis and visualization of functional magnetic resonance neuroimages. Comput. Biomed. Res. 29, 162 -- 173 (1996). 53. Cole, M. W., Bassett, D. S., Power, J. D., Braver, T. S. & Petersen, S. E. Intrinsic and Task-Evoked Network Architectures of the Human Brain. 83, 238 -- 251 (2014). 54. Siegel, J. S., Mitra, A., Laumann, T. O. & Seitzman, B. A. Data quality influences observed links between functional connectivity and behavior. Cerebral … (2016). 45. Shine, J. M. et al. The Dynamics of Functional Brain Networks: Integrated 46. Lerman, C. et al. Large-scale brain network coupling predicts acute nicotine 55. Gu, S., Satterthwaite, T. D. & Medaglia, J. D. Emergence of system roles in 56. Zalesky, A., Fornito, A. & Bullmore, E. On the use of correlation as a measure of 57. Rosvall, M., Rosvall, M., Bergstrom, C. T. & Bergstrom, C. T. Maps of random 58. Brandes, U., Delling, D. & Gaertler, M. On modularity clustering. IEEE 59. Lancichinetti, A. & Fortunato, S. Community detection algorithms: A comparative 60. Berenstein, A. J., Piñero, J., Furlong, L. I. & Chernomoretz, A. Mining the Modular 61. Clauset, A., Newman, M. & Moore, C. Finding community structure in very large 62. Blondel, V. D., Guillaume, J.-L., Lambiotte, R. & Lefebvre, E. Fast unfolding of normative neurodevelopment. in (2015). network connectivity. NeuroImage 60, 2096 -- 2106 (2012). walks on complex networks reveal community structure. 105, 1118 -- 1123 (2008). transactions on … (2008). analysis. Phys. Rev. E 80, 056117 -- 11 (2009). Structure of Protein Interaction Networks. PLOS ONE 10, e0122477 (2015). networks. Phys. Rev. E (2004). communities in large networks. Journal of Statistical Mechanics: Theory and Experiment 2008, P10008 (2008). https://www.humanconnectome.org/storage/app/media/documentation/s1200/HC P_S1200_Release_Reference_Manual.pdf 64. Shen, X. et al. Using connectome-based predictive modeling to predict individual behavior from brain connectivity. Nat Protoc 12, 506 -- 518 (2017). 63. WU-Minn HCP 1200 Subjects Data Release Reference Manual. 1 -- 213 (2018).
1712.02142
1
1712
2017-12-06T11:46:02
Maps of Visual Importance
[ "q-bio.NC" ]
The importance of an element in a visual stimulus is commonly associated with the fixations during a free-viewing task. We argue that fixations are not always correlated with attention or awareness of visual objects. We suggest to filter the fixations recorded during exploration of the image based on the fixations recorded during recalling the image against a neutral background. This idea exploits that eye movements are a spatial index into the memory of a visual stimulus. We perform an experiment in which we record the eye movements of 30 observers during the presentation and recollection of 100 images. The locations of fixations during recall are only qualitatively related to the fixations during exploration. We develop a deformation mapping technique to align the fixations from recall with the fixation during exploration. This allows filtering the fixations based on proximity and a threshold on proximity provides a convenient slider to control the amount of filtering. Analyzing the spatial histograms resulting from the filtering procedure as well as the set of removed fixations shows that certain types of scene elements, which could be considered irrelevant, are removed. In this sense, they provide a measure of importance of visual elements for human observers.
q-bio.NC
q-bio
Maps of Visual Importance Xi Wanga,∗, Marc Alexaa aTechnische Universitat Berlin, Germany Abstract The importance of an element in a visual stimulus is commonly associated with the fixations during a free-viewing task. We argue that fixations are not always correlated with attention or awareness of visual objects. We suggest to filter the fixations recorded during exploration of the image based on the fixations recorded during recalling the image against a neutral background. This idea exploits that eye movements are a spatial index into the memory of a visual stimulus. We perform an experiment in which we record the eye movements of 30 observers during the presentation and recollection of 100 images. The locations of fixations during recall are only qualitatively related to the fixations during exploration. We develop a deformation mapping technique to align the fixations from recall with the fixation during exploration. This allows filtering the fixations based on proximity and a threshold on proximity provides a convenient slider to control the amount of filtering. Analyzing the spatial histograms resulting from the filtering procedure as well as the set of removed fixations shows that certain types of scene elements, which could be considered irrelevant, are removed. In this sense, they provide a measure of importance of visual elements for human observers. Keywords: Visual importance, Mental imagery, Eye tracking 1. Introduction The relative importance of visual elements presented on a display is useful in a variety of applications. In the fields of visualization and human-computer interaction, it is often a design goal to make specific elements clearly visible. Consider for example ∗Corresponding author at Computer Graphics, Technische Universitat Berlin, Germany. Email address: [email protected] (Xi Wang) 1 the visualization of medical imaging data: it is important that anomalies are recognized by the physician inspecting the data. Likewise, in computer graphics or image and video coding, it is of interest what visual objects are important or unimportant in order to allocate resources such as computation time (for example, for rendering 3D objects in an augmented reality scenario) or bandwidth (as in video coding). Note that in these applications we are not just interested in whether visual elements have been processed by an observer, but we need a map of the relative importance of regions of the stimulus. The common approach for creating such maps is based on eye tracking and accumulating fixations from several observers (Bruce and Tsotsos, 2005; Judd et al., 2009, 2012; Coutrot and Guyader, 2014; Xu et al., 2014; Borji and Itti, 2015) The underlying reasoning is that most visual information is processed during fixating (Matin, 1974), and saccade target selection is likely optimized to attend to the most important visual elements first (Itti and Koch, 2000; Parkhurst et al., 2002). A va- riety of factors have been shown to affect the selection (Krieger et al., 2000; Einhauser et al., 2008; Kienzle et al., 2009; Tatler et al., 2011; Schomaker et al., 2017). We are nonetheless unconvinced that fixations alone are sufficient for creating maps of importance of visual elements, for several reasons: fixated locations are not neces- sarily attended; and the frequency and duration of fixations is not always correlated with the importance of an object as assigned by an observer. Here, we suggest a new experimental setup that creates a map based on whether visual elements have been encoded and deemed important. 1.1. Fixations vs. visual attention It is well known that observers often fail to detect seemingly prominent visual ob- jects, typically because they are given a task for which the object is irrelevant – an effect called inattentional blindness (Simons, 2000). One may speculate that the specific task leads to a lack of fixations on the unexpected object, yet this is not the case. In partic- ular, Memmert (2006) repeated the famous Gorilla experiment while tracking the gaze of the observers. They found that observers fixate the Gorilla a significant amount of time and that recognizing the Gorilla was independent of the fixations. Koivisto et al. (2004) used a more controlled setup with a screen presentation of the stimulus. Their 2 particular findings are that visual attention and fixations were uncorrelated under the specific experimental conditions (they note, however, that this might not be common under natural viewing conditions). 1.2. Fixation statistics vs. importance A variety of factors influence the statistics of fixations, some of which may be un- correlated with the importance of a scene element that is later attributed by an observer. To begin with, a confounding factor is the experimental setup for gathering fixa- tions: it is common to only consider fixations in the first few seconds after the onset of the stimuli. This is based on the intuitive idea that the most important elements would attract visual attention first. Indeed, it has been shown that fixation locations relate less to image statistics over time (Itti and Koch, 2000; Parkhurst et al., 2002). This approach, on the other hand, limits the number of locations extracted, as each fixation has a certain minimal duration (typically not less than 150ms (Pannasch et al., 2008)). Also the distance between fixations (the saccade amplitude) is limited to about 7◦ vi- sual angle. These characteristics of fixations, saccades, and the time in which fixations are collected have an important consequence: only few elements can be visited and, since the distribution of fixations is constrained but not the distribution of scene ele- ments, it is not clear that all fixations indeed correspond to visually important elements. The common strategy for coping with this potential problem is to average over the data of repeated trials and several observers (e.g. Van Der Linde et al., 2009; Ramanathan et al., 2010; Koehler et al., 2014). It remains unclear why averaging would result in a measure of importance. More importantly, it is known that fixations may extend or accumulate for a variety of reasons, such as generally being biased towards the center of the field of view (Tatler, 2007), being longer for scene elements that are unlikely or improbable given the scene context (Loftus and Mackworth, 1978; De Graef et al., 1990), or following the gaze of humans depicted in the scene (Ricciardelli et al., 2002; Hutton and Nolte, 2011). We also speculate that visual elements that are complicated (e.g., cluttered or obstructed) require longer or more fixations. In all of these cases, the importance of the visual element depends on a variety of factors, and assigning high importance based on just 3 Figure 1: An example of mirror scene. the number or duration of fixations is unwarranted. As an example for this statement consider a mirror in the scene (see Figure 1): the mirror likely draws fixations because it reflects a variety of interesting objects. We would suspect, once the region is being decoded as a mirror it likely decreases in importance, either because the objects are visible in the scene and can be attended to directly, or because understanding the scene through a mirror is too complicated to be attempted during a brief exposure to the stimulus. 1.3. Importance of fixation based on recall In this work, we distinguish among fixated regions that do and do not make it into the visual memory; and among the ones being stored we create a map that reflects their relative importance. The contribution of the paper is an experimental setup and an algorithm for post-processing that leads to such maps. The idea is to base importance on the motion of the eyes during the recall of the image. It is know that the eyes move during recalling an image. In the absence of other visual features (i.e. while looking at nothing), the motion of the eyes is related to the motion of the eyes while the stimulus was present. In particular, fixations during recall are related to fixations during exposure with the stimulus (Johansson and Johansson, 2014; Laeng et al., 2014). We want to use fixations during recall as indicator of the presence in visual memory, and by averaging over multiple observer as a measure for 4 their importance. The fixations during recalling an image while looking at nothing lack a reference frame and cannot be used directly to identify important elements in a visual stimulus. Our idea is to record the exploration of an image using a calibrated eye tracker as well as the immediate recall of the image. The fixations during exploration of the present stimulus can be mapped to locations in the image. And the fixations during recall can be used to validate the fixations during exploration. Despite the large body of work on eye motions during recall while looking at noth- ing, we were unable to find data from experiments. In order to analyze our idea, we have performed an experiment based on standard image data, in which we record eye movements during exploration and recall of those images. The complete data set for pairs of exploration and recall sequences will be made publicly available. 1.4. Eye movements during recall The looking at nothing phenomenon, i.e. the effect of eye movements during the recall of a scene while looking at a neutral visual stimulus, has been known for a long time (Moore, 1903), yet was mostly considered an odd coincidence. Laeng and Teodor- escu (2002) are usually credited with having established that eye movements during re- call play a functional role in memory retrieval. Their experiment is based on inhibiting eye motion (here, by asking observers to fixate on a point) during either exploration or recall. Inhibiting eye motion during exploration lead to reduced eye motion during recall, and inhibiting eye motion during recall let to decreased recall performance. A variety of other setups have confirmed and refined these findings (Johansson and Jo- hansson, 2014; Laeng et al., 2014; Scholz et al., 2015; Bochynska and Laeng, 2015; Pathman and Ghetti, 2015; Scholz et al., 2016). In particular, it has been shown that the fixations during recall of the stimulus are connected to the location of the objects. Ferreira et al. (2008) noted that observers' gaze would move to the location of previously observed visual elements when cued by keywords, even if the visual element was not present anymore. de Vito et al. (2014) confirmed that inhibiting eye motion during recall decreased memory performance, however, they refined the results and showed that only memory of spatial location 5 was affected, while recollection of visual content was independent of eye movements. Martarelli et al. (2017) showed that gaze still indicates spatial location if observers were asked to imagine a scene with objects being replaced by other similar objects. All these experiments provide evidence that the eye position during recall of a stimulus reveals the location of an object. 1.5. Accounting for displacements in recall data We want to use the fixations during recall to validate the fixations during explo- ration. The problem in this task is that the fixations in the recall phase are only qual- itatively related to the stimulus – without the stimulus being present there is no frame of reference. We also find that the motion of the eye during recall contains significant local displacement, i.e. the spatial encoding in eye position contains error. We develop an approach for the approximate matching between the fixations in the recall sequence and those in the corresponding exploration sequence. It is based on computing a de- formation mapping for the locations of fixations in the recall phase. After applying the deformation we retain fixations from the exploration sequence that are close enough to a fixation in the mapped recall. A threshold on the distance between deformed fixa- tions in the recall and fixations in the exploration sequence allows us to steer the filter. This leads to varying maps of importance, going from considering all fixations in the original exploration sequences to subsets. We explore that resulting spatial importance maps and observe a variety of meaningful effects. 2. Method We collect eye movement data during the exploration and recall of images. We first describe how we acquire this data and motivate our design choices for the experiment. To our knowledge, this is the first experiment in this direction. We then provide statis- tics of the acquired data and show that it is consistent with other data collected for the free-viewing task. 6 2.1. Experiment The basic setup for our experiment is to present observers with an image as stim- ulus and let them recall the image immediately after the presentation of the stimulus. Their head remains in the same position during stimulus presentation and recall. In the following we discuss the details. 2.1.1. Participants We recruited 30 participants for our experiment (mean age = 26, SD = 4, 9 female). Importantly, all participants are All reported normal or corrected-to-normal vision. naive w.r.t. the aim of using the fixations during the recall phase to gauge the impor- tance of fixations during exploration. All studies have been carried out in accordance with the Code of Ethics of the World Medical Association (Declaration of Helsinki). Participants were informed about the procedure before giving their written consent and could stop the experiment at any time. Their time was compensated and all data is used anonymously. 2.1.2. Apparatus The experiment was conducted in a dark and quiet room. A 24-inch display (0.52m× 0.32m) with a resolution of 1920×1200 pixels was in front of the observer at a distance of 0.7m. We used an EyeLink1000 desktop mount system (SR Research, Canada) to record the eye movements at a sampling rate of 1000Hz. A chin and forehead rest was used for stabilization. All experiments were conducted in binocular viewing condition, but only the movements of the dominant eye were recorded. 2.1.3. Stimuli We use 100 natural images randomly selected from the MIT data set (Judd et al., 2012). This set includes both indoor and outdoor scenes of various complexity (see Fig- ure 4 for example images). These images have been used in eye tracking experiments with a free viewing task and the data is publicly available (Judd et al., 2012). This allows us to compare the eye tracking data we gathered during the exploration phase to existing data. All images were presented at the center of display in their original size with the largest dimension being 1024 pixels. 7 2.1.4. Procedure The experiment consists of 100 basic trials, and a single image was presented in each trial. The details of the presentation in one trial are: Prior to the presentation of the image, the screen is black and shows a white dot in the center (1◦ visual angle). Observers are asked to fixate at the white dot so that the eye motion starts in a consistent way for all images and observers. Then the image is presented for 5 seconds. Observers are asked to explore the image in order to later be able to recognize it (a free viewing task). After presentation of the image white noise is shown for 0.5 seconds to suppress the after image. Then the screen is set to neutral gray for 5 seconds. During this phase, observers are asked to immediately recall the image from memory. After that the screen turns black for 1.5 seconds before the procedure is repeated for the next image. Each observer saw all 100 stimuli and the order of them is randomized for each participant. Instructions as to encourage observers to explore the mental image during the recall phase were given at the beginning of the experiment. The 100 trials were divided into five blocks. Each block of 20 trials started with a standard 9-point calibration proce- dure. We repeated the calibration until the average accuracy reported in the following validation was below 0.5◦ and no validation point had an error larger than 1.0◦. After a successful calibration 20 trials were performed. This procedure required roughly 5 minutes. All participants performed the experiment, however, for two of them we were unable to achieve the desired calibration accuracy (probably due to glasses they had to use). This leaves us with data sets from 28 observers. An important design choice was the duration of the exploration and recall peri- ods. In particular the recall phase cannot be too short, because retrieving the image from memory requires time. We decided to follow experiments in cognitive psychol- ogy (Laeng et al., 2014) and allow 5 seconds for recall. For reasons of temporal equiv- alence, we also present the stimulus for 5 seconds. The basic task of recalling an image turns out to be significantly more demanding than just exploring the image. To avoid exhaustion we limited the number of presentations to 100 while still having a suffi- cient number of repeats for our analysis. Participants were allowed to take a break of arbitrary length after each block. 8 The whole experiment lasted about one hour. At the end of the experiment partic- ipants were shown 10 images, half of which were part of the 100 stimuli used in the experiment. Images were presented one after the other in a randomized order, and par- ticipants had to decide if the images were among the 100 presented to them. All parts of the experiment were explained in detail at the beginning, including the memory test at the end. 2.2. Data processing and analysis We congregated and analyzed the data from 28 observers for the exploration and recall phase of each image. 2.2.1. Processing raw data and fixation detection We base our analysis on fixations. To extract them from the raw eye tracking data, we use a dispersion based algorithm (Holmqvist et al., 2011), which involves setting two thresholds: minimum duration τ and maximum dispersion ϕ. Setting these pa- rameters is a standard procedure for eye-tracking experiments with the stimulus being present, we feel experimentation is warranted for the recall phase. We tested a range of parameter combinations for values discussed in the literature and Figure 2 shows the classification results of one sequence under different parameter settings. We found that also during recall the settings were not critical and within the common range detected fixations did not vary much. Thus we fixed the parameters for the following analysis. The minimum duration of fixation is set to be 100 ms and maximum dispersion is set to be 0.5◦ visual angle. We detect fixations in all exploration and recall sequences. During exploration the median and mean number of fixations is 16 (SD=2.8), and during recall the median and mean number of fixations is 11 (SD=3.6). The fewer fixations in recall have a correspondingly longer duration (avg = 452.2 ms, SD = 308.0 ms) than fixations in exploration (avg = 278.0 ms, SD=73.4 ms). Fixation durations are plotted as a function of the starting time of the fixation during the trial in Figure 3. Exploration phase is separated from the recall phase. Each fixation in the dataset from all 28 observers is depicted as a cross and the black curve shows the averaged duration in each phase. 9 Figure 2: Fixation classification results for different parameter settings. The minimum duration threshold τ ranges from 50 ms to 200 ms and the maximum dispersion threshold ϕ ranges from 0.25◦ to 1.0◦. Raw samples are depicted in green and fixations in red whose radius correspond to the durations. Fixation number (#fix) and mean duration (md) are reported in each plot. As shown in the first row, small dispersion ϕ results in more fixations with shorter durations and a large ϕ tends to fuse fixations together. Second row depicts the influence of τ, the minimum duration threshold. Small τ leads to over segmentation and large τ results in much less fixations. After the first two fixations, there is no significant variation of fixation duration during the trials for both exploration and recall (the durations getting shorter towards the end of the trial is an artifact of stopping data collection after 5 seconds). On average fixations in recall are roughly two times longer than the ones in exploration. The fixations from all participants for a single image can be summarized in a spatial histogram, leading to a so-called heat map for the image. We do this for the data from our experiment and the publicly available data for the 100 images. Similar to Judd et al. (2012) we remove the first center fixation from each sequence and apply a Gaussian filter with a kernel size equivalent to 1 degree of visual angle. 2.2.2. Data from exploration As a sanity check, we compare the heat maps resulting from the publicly available data to the heat maps from the exploration phases in our experiment. To measure the similarity of corresponding maps we use Pearson's correlation coefficient (CC), 10 Figure 3: Comparison of fixation durations in exploration (a) and recall (b). Fixation durations are plotted as a function of starting time . All fixations from 28 observers for 100 trials are depicted as crosses. The black curves indicate the mean durations and the center 50 percent intervals are depicted in blue. following Bylinskii et al. (2016). Averaging over all 100 pairs of saliency maps, we find a mean CC of 0.766 with standard deviation 0.115. We also perform a qualitative comparison. Figure 4 shows maps for different ranks among the correlations. Most heat maps are very similar. For those that are dissimilar, we see a stronger center bias in the existing data. We speculate that this is due to the shorter presentation time of 3 seconds vs. the 5 seconds in our experiment. We conclude that our results are qualitatively similar to previous experiments with the same data. 2.2.3. Data from recall Compared to exploration sequences, recall sequences have fewer but longer fixa- tions (see. Figure 3) The last row in Figure 4 shows the heat maps generated from the recall phases. As can be seen, while they roughly resemble the maps from the ex- ploration phase, they are more biased towards the center and typically fail to exactly correspond with features in the image. In particular, they cannot be used directly as heat maps for the images. In some cases, the correspondence between fixations in exploration and recall is clear. Then we found that the temporal order is generally not preserved (see Figure 5 for some examples). This is consistent with previous results in cognitive psychol- ogy (Johansson et al., 2012). 11 Figure 4: Comparison of heat maps. Images are ranked based on correlation scores between the heat maps generated using the existing data from Judd et al. (2012) and the data we collected during presentation of the images. Seven examples are provided. Each column shows the original image, the MIT heat map, and the heat maps created from the exploration and recall phases of our experiment. Figure 5: Pairs of exploration and recall fixation sequences from four observers. First row shows fixation sequences collected during exploration and second row shows the corresponding recall sequences. The temporal order of each sequence is indicated by the numbers and consecutive fixations are connected by lines. We also notice that observers tend to 'stall' during recall. This means they stop moving their eyes, leading to fixations that are unlikely corresponding to image con- tent. This sometimes happens for long durations at the center. For 5 out of the 28 participants, the number of fixations in recall is less than half compared to exploration. One participant reported after the experiment that he changed his strategy through the 12 ImageMIT Heat mapViewingHeat mapcc = 0.448cc = 0.662cc = 0.748cc = 0.926cc = 0.858cc = 0.838cc = 0.789rank 1rank 17rank 34rank 50rank 67rank 83rank 100RecallHeat map Figure 6: Subset selection criteria. We base our selection on the similarity between the averaged exploration and recall heat maps. Data sets are ranked according to correlation scores. Two different types of inconsis- tencies are visible in the least similar pairs: the distributions of recall fixations are either peaked at one point or spread more randomly. experiment and only recalled the single most interesting element of the image. 2.3. Selecting a subset of observers The analysis above suggests that not all observers consistently moved their eyes to revisit important features in the image. Without ground truth it is impossible to decide if the data from a particular observer is useful. In order to avoid basing the analysis on low quality data, we focus our analysis on only half of the observers. For selection, we take the fixations for all 100 images from one observer and com- pare the spatial distribution of fixations during exploration and recall. Figure 6 shows heat maps for several observers. The similarity of the spatial distributions varies; some show significantly different patterns (see the three examples on the right hand side in Figure 6). The inconsistencies are of two types: 1) the spatial distribution in recall has smaller variance compared to exploration (indicating observers fixated mostly at one location), 2) the spatial distribution has much higher variance. Both types of differ- ences hamper further using the fixations from recall for matching against the fixations in exploration. To account for the idiosyncrasies of individual observers we compare the spatial distribution from exploration to the spatial distribution from recall for each observer (rather than judging only the spatial distribution during recall). As an unbiased method to measure their similarity we compute Pearson's correlation coefficient. This results in a ranking of the observers - the images in Figure 6 are sorted based on this ranking. We select the upper half of this ranking for the following processing and analysis. The 13 Figure 7: Example of the location mismatch between fixations in exploration and recall. Fixations from exploration are shown in green (a); recall fixations are subset of the fixations during exploration and are shown in red. Their displacement (b) contains global rigid transformation (c) and deformation (d). The deformation mapping is based on the yellow consensus locations (e). publicly available data will include all observers. 3. Filtering the fixations based on recall witnesses Our idea for exploiting the information in the recall is to match the fixations during exploration with the fixations during recall. A fixation in recall matched to a fixation in exploration serves as a witness, testifying that the visual element corresponding to the fixation is important. 3.1. Deformation mapping The problem of establishing the matching is that spatial locations of the fixations in recall are distorted relative to the locations of features in the image. This distor- tion could be decomposed into a global rigid transformation (due to a lack of reference frame) and local inaccuracies. Figure 7 provides an illustration: The six points in a) depict a sequence of fixations during exploration, of which only five are being re- called. The observed set of fixations during recall projected into the same reference frame is shown in b). We may think of the locations as being composed of the rigid transformation in c) and the deformation in d). The deformation contains no global rigid transformation in the sense that moving or rotating the fixations would not make the sum of the squared distances to the matched fixations in exploration any smaller. Note that fixations in recall and exploration that are corresponding cannot be found by simply considering their distances. 14 When data is corrupted by global transformation and deformation, it is common to approximate the effects by minimizing the squared distances between matched data points. Yet, this is a chicken and egg problem: The estimated mapping is supposed to improve the matching, while the matching is needed to estimate the mapping. The common solution to this dilemma is to start with a first guess about the matching, then compute the deformation, and based on the deformation make a better guess about the matching, and so on until the method converges. A common approach of this type is Iterated Closest Point (ICP) (Besl and McKay, 1992), which is used to compute a global rigid transformation to match two partially overlapping point sets. The approach we suggest for matching the recall fixations to the fixations in ex- ploration is inspired by ICP, yet extends it in two important ways: First, rather than using a closest point matching, it matches fixations in recall to consensus locations in the exploration data. Second, rather than computing a global rigid transformation it is based on a deformation mapping that has controllable overall deviation from a global rigid transformation. The first modification is motivated by the situation in Figure 7 b), namely that matching the recall fixations to the closest fixations in exploration may lead to wrong assignments. These wrong assignments will steer the estimated mapping away from the desired solution. Given the data, we believe several situations have to be accom- modated: 1. One fixation in recall maps to exactly one fixation from exploration. 2. One fixation in recall maps to several close fixations from exploration, i.e. the observer recalls just one scene element that drew several fixations. 3. A fixation in recall may have no corresponding fixation in exploration (i.e. be- cause the fixation is unrelated to the process of recalling, or the spatial error is too large to be rectified). We suggest to accommodate all cases by computing a consensus location for each recall fixation, based on computing a weighted average of the fixation locations in exploration. The weights decay exponentially with distance. This has the effect that if only one fixation in the exploration sequence is close while the others are far away 15 (case 1) the closest location will receive a large weight, while the others relatively small weights, so that the consensus location will be the matching fixation. In case several fixations in exploration are close (case 2), all of them receive equal weights, and the consensus location is the center of these fixations. If no fixation is found for matching (case 3), all fixations receive little weight, and the consensus location is roughly where the recall fixation already is. Figure 7 e) shows the consensus locations for the situation in b). The computation of consensus locations is explained in Appendix A. It can be controlled by a parameter wp, measured in visual angle, which could be interpreted as distance of fixations that contribute significantly to the weighted averaging procedure. We set the parameter to wp = 2◦ and discuss this choice in Section 6. The consensus locations for recall fixations are used to compute a deformation mapping D : R2 (cid:55)→ R2 from the current positions of the recall fixations to the desired ones. Allowing deformation overcomes the problem that the best we could achieve with computing a rigid transformation is to be left with the deformation, i.e. the situa- tion depicted in Figure 7 c). However, some global rigidity needs to be preserved, i.e. the positions should not deform arbitrarily. This is important, as we would otherwise always match all recall fixations to some fixations in exploration. For example, if we allowed arbitrary scale, it would always be possible to scale the set of recall fixations to a single point and then match it to one of the fixations in exploration. To avoid such degenerated solutions it is important to restrict the mapping to preserve the global structure. The computation of the mapping based on consensus locations and the cur- rent positions of the recall fixations is explained in Appendix A. The global rigidity of the mapping can be controlled by a parameter wd, measured in visual angle. Roughly speaking, wd describes the distance of points that may be transformed by two rigid transformations that differ significantly. If wd is large, fixations are transformed by very similar rigid transformations, restricting the deformation; if it is small, fixations are transformed by independent rigid transformations, allowing deformation. We set the parameter to wd = 10◦ and discuss this choice in Section 6. The necessary steps to compute the deformation mapping are given in pseudocode in Appendix A. Once the mapping is computed, we simply use distance as the sole criterion for testimony: a fixation in exploration has a witness, if there is a mapped 16 Figure 8: Quantitative results of filtering relative to witness radius  ∈ [1◦, 10◦]. Results from sequences in the selected subset of observers are shown in orange and the results from the the excluded subset are shown in blue. (a) Relative number of fixations retained after filtering. Each observer's data is scattered in the graph and the light color marks the center 50% region of the distribution. The effect of the witness radius is uniform across different images and observers. (b) Percentage of all filtered maps whose most salient spots remain within 4◦ after filtering. fixation in recall that is closer than the witness radius . 4. Quantitative results of filtering Filtering is supposed to reduce the number of fixations in a meaningful way. It is affected by the choice of witness radius . Before we analyze specific effects on a qualitative level we present global statistics to show that the effect of filtering is not random. One statistic is the number of fixations that remain after filtering. We expect that even for relatively small witness radii a significant number of fixations remain. The left graph in Figure 8 depicts the number of retained fixations relative to the witness radius. Here, we differentiate between the subset of observers we selected for further processing vs. the subset of excluded observers. Note that the effect of the threshold is quite similar across different images and observers. For values larger than  = 10◦ we effectively retain all fixations. For very small values the amount of noise in the recall data leads to results that seem to contain no useful information. In the subset of observers we consider more reliable fixations remain after filtering. This means 17 that more fixations in the exploration sequences found a closest match in the recall sequences. This is what we expect, as the distributions of the fixations in exploration and recall are similar. We also analyze if the sets of fixations change in the sense of their relative impor- tance within an image. For this we identify in each image the region with the most fixations during exploration based on the smoothed spatial histogram. We do the same after filtering and consider that the point has shifted if the difference in spatial location exceeds 4◦. The right graph in Figure 8 shows the result relative to the witness radius. We observe that, as expected, for large witness radii there is no significant change, yet for smaller radii the most salient object changes in about 40% of the images. We take this as indication that filtering is not random: on one hand, we expect filtering to affect the relative importance of elements in the scene; on the other hand, in a significant number of images the most fixated object is indeed important. Figure 9 illustrates the effect of varying the threshold on the resulting maps for two of the images using the selected subset. We show heat maps for the values  = 1◦, 2◦, 4◦ and observe different effects: some regions that attracted a lot of fixations appear to have been deemed relatively unimportant, for example the sign to the left in the top image. Other regions are boosted in importance after filtering such as the objects in the foreground in the top image or the tower to the right of the road in the bottom image. In the following section we attempt a systematic discussion of these effects based on the selected data. 5. Main results We apply the deformation process to each pair of fixation sequences for all images from the subset of selected observers. Then we apply the recall based filtering based on witness radius  = 1◦. In other words, a fixation is retained only if there is a corresponding fixation in the deformed recall sequence that is within 1◦ visual angle. This is a rather strict setting, removing a lot of fixations. We chose it for achieving a significant effect, at the expense of possibly introducing more noise. Our analysis is based on comparing the heat maps from the unfiltered and filtered 18 Figure 9: Filtered maps using different thresholds. We show examples of  = 1◦, 2◦, 4◦. Original heat map shows the result based on all fixations during exploration of the image. The color coding of the scaled maps is based on the highest value appearing in all heat maps, which is naturally the one generated without filtering. The normalized map is color coded in its own range. images averaged over the selected 14 observers. For each image, we detect the region showing the largest (absolute) difference in the heat maps (these difference maps are also shown in all illustrations). The inspection of these 100 difference maps showed that the removal of fixations from the exploration sequences can be explained in most cases based on established effects. Before we exemplify these effects, we present an explanation for the removal of a fixation that appears to apply to virtually all cases. For a region to be removed it necessarily received many fixations during exploration but few during recall. Indeed, most of these regions show complex visual patterns, which explains why they have received a lot of fixations during exploration. Then one of two cases seems to apply: 1) despite the effort, the region could not be decoded into a semantic object and is, consequently, not stored; 2) an object has been recognized, but is deemed unimportant (relative to the rest of the scene, or at least relative to the 19 Image Stimulus ε = 1°ε = 2° ε = 4°Original Heat MapScaled by Original Heat MapNormalized Normalized Image Stimulus ε = 1°ε = 2° ε = 4° ε = 1°ε = 2° ε = 4°Original Heat Map ε = 1°ε = 2° ε = 4°Scaled by Original Heat Map small amount of time allowed for recalling). In the following we discuss mostly effects of the second type. Clearly, these effects may vary among observers – in particular, the importance of a recognized object is highly subjective. For effects in the whole dataset of 100 images see the supplementary material. 5.1. Low-level features Low-level features have been shown to contribute to saccade target selection (Bad- cock et al., 1996; Krieger et al., 2000; Kienzle et al., 2009). Consequently, they con- tribute significantly to fixation-based attentional models. In our model that considers also attention, awareness, and relative importance of the visual object, their strength is no more relevant for the relative importance of a visual element. Figure 10 shows three examples of such cases. A white box in the right corner over a black background in the first image draws a lot of fixations due to its high contrast. It is less dominant in the filtered heat map as shown in both the third and last columns. We would speculate that observers either were unable to decode the object or decide that a white box-like object is not important for the scene. Similar effects can be observed in the second example on the red car (which is hard to be recognized as such) and in the third example on the ball (which is relatively unimportant compared to its colorfulness and contrast). Instead we see fixations on the train and the shape of the lamp become relatively more important. 5.2. High-level features Among high-level features, we observe noticeable effects on text, signs, and people. 5.2.1. Text and signs Text and signs commonly draw attention as they provide a lot of information in small spatial region. Some signs are purposefully designed to be conspicuous. As shown in Figure 11, they have relative high attendance during exploration. Their im- portance depends significantly on the result of decoding their meaning and then on the experimental conditions, for example the task, but also the experience of the observer. Based on our selection of examples, the elements in Figure 11 significantly decreased 20 Figure 10: Strong low-level features after filtering. The second column shows the heat map based on all fixations during exploration of the image. Removed fixations are used to generate the removal map (or difference map) and remaining fixations are used to generate the filtered heat map, which is essentially the difference between original heat map and the removal map. The color coding of the removal map is scaled by the highest value in original heat map, while the filtered heat map is normalized in its own range. Fixations triggered by low-level features are largely removed in the filtered heat maps, if the underlying visual object was difficult to understand or turned out to be of little relevance. in importance. We believe the text in the first row and the sign in the last row were undecipherable, while in the other two examples the text (whether legible or not) and the exit sign were deemed unimportant for the scene. 5.2.2. People As shown in Figure 12, observers look at humans in images regardless of their size. The relative importance of visual objects recognized as people apparently depends on the scene. In the two examples above the dashed line, people are not a significant element of the scene – they are expected and not the dominating elements. In the other two examples, humans are contributing more to the composition of the scene, and they remain important after filtering. 21 Image StimuliOriginal Saliency MapNormalized Filtered Saliency MapRemoval Map Figure 11: The effect of filtering on text and signs. The meaning of the color coded images is as before. Illegible or unimportant text (upper rows) as well as signs without particular importance for the scene (lower rows) lose importance. 5.3. Relative importance of similar items Similar objects in a scene may be assigned equal or different importance, indepen- dent of their visual representation. We provide several examples in Figure 13. Similar objects may well be recalled similarly, as is demonstrated in the first example. In the second row, the larger boat dominates the fixations. Based on recall, the smaller boat in the middle row gains in relative importance. Among a set of faces, all of which draw fixations, the one that stands out indeed becomes more important. 5.4. Other effects When looking at an image we often follow the apparent gaze of people in the scene because it might provide additional information about people's intention or the local 22 Image StimuliOriginal Saliency MapNormalized Filtered Saliency MapRemoval Map Figure 12: Filtering may or may not affect people. The meaning of the color coded images is as before. People in the scene are observed in exploration but their importance in the filtered heat maps depends on the scene composition. In the upper rows, the importance of people decreased, while in the lower rows they remain important. environment (Corkum and Moore, 1998; Hutton and Nolte, 2011; Gallup et al., 2012; Bayliss et al., 2013). This effect is notoriously hard to model computationally (Xu et al., 2014; Recasens et al., 2015). In our model, the object found in these locations determines the importance of the scene element, and not the mere fact that some person in the scene appears to be looking at the region (see Figure 14). Other noteworthy visual elements are horizons and mirrors. Both attract a lot of fixations, however, appear of little importance when the image is being recalled (see Figure 15). 23 Image StimuliOriginal Saliency MapNormalized Filtered Saliency MapRemoval Map Figure 13: Filtering results for elements with similar meaning. The meaning of the color coded images is as before. In the top row, similar items preserve their similar importance. The smaller boat in the middle row gains in relative importance, while still being dominated by the larger boat. In the last row, the face that stands out among other faces increases in importance. Figure 14: The point of gaze of people in the stimulus is effectively filtered. The meaning of the color coded images is as before. In our recall based attention model, the apparent gaze seems to have less impact and the resulting heat map rather depends on the scene elements being gazed at. 24 Image StimuliOriginal Saliency MapNormalized Filtered Saliency MapRemoval MapImage StimuliOriginal Saliency MapNormalized Filtered Saliency MapRemoval Map Figure 15: Other effects in the filtered saliency maps. The meaning of the color coded images is as before. Observers scan the horizon, yet only recall elements if they were dominant. Likewise, mirrors are complex visual stimuli, but once they are decoded they become unimportant. 6. Discussion 6.1. Data collection While the fixations we extracted from the collected eye tracking data in the explo- ration phase are similar to the reference (Judd et al., 2012), there are also some images with considerably different resulting heat maps. Our experimental conditions differ only in the length of exposure, namely 3s vs. 5s in our setup. As the data is used for judging the success of computational models for the prediction of saccade targets, we wonder about the variation in the data across observers, and minor variation in experi- mental conditions, relative to the performance difference of different models. We noticed that recall is a strenuous task and not all observers are willing or able to continuously move their eyes. It remains unclear if a lack of motion means that observers paused the recall, however, the fact that the recall performance correlated with motion of the eyes in other experiments (Johansson and Johansson, 2014; Laeng et al., 2014; de Vito et al., 2014) suggests that this is the case. It is unclear how to motivate observers to engage more actively in recall. Observers only had limited time for recall, thus limiting the number of elements that could potentially being recalled. This situation is similar to the influence of time for recording the fixations during exploration: shorter times leads to fewer fixations and more variation across observers, while longer times increases the amount of noise (Itti 25 Image StimuliOriginal Saliency MapNormalized Filtered Saliency MapRemoval Map and Koch, 2000; Parkhurst et al., 2002). It is not clear how our choice of 5 seconds for recall affected the results. If it was acceptable to present the stimulus for a significantly longer time, a sub- stantially different protocol for collecting information could be used: observers use standard interaction devices such as computer mice or touch screens to mark inter- esting points or objects (Jiang et al., 2015; Kim et al., 2015, 2017). This higher-level interaction leads to selection based on semantics of features. It has been recently shown to be inconsistent with fixations (Tavakoli et al., 2017). 6.2. Subset selection We selected only half of the observers based on the correlation of the spatial his- tograms between exploration and recall. The choice of 14 is arbitrary and it appears that varying it has almost no effect on the outcome. However, we do observe that the data of few observers is different and would have obscured the results we show. For ranking observers we also considered using the pair correlation function (Illian et al., 2008; Oztireli and Gross, 2012) as it has been recently demonstrated to char- acterize the spatial distribution of fixations well (Trukenbrod et al., 2017). Yet, we felt there is no obvious way to compare the distributions. Depending on the choice of parameters and method for measuring the similarity the results varied, while largely re- sembling the order computed with the much simpler approach of taking the correlation. Consequently, we opted for correlation for the subset selection. 6.3. Deformation mapping As the true matching is unknown, it is impossible to optimize parameters control- ling the matching and deformation automatically. We had asked 3 participants in the experiment to manually mark the correspondences among the fixations, in this way generating a reference, yet their data are unreliable (consistent with the literature (Vo et al., 2016)). On the other hand, the results of the deformation process are stable across a wide range of parameters (see Figure 16). We found that the radius for matching recall fixations to exploration fixations wp should be chosen in the range of 2◦ − 4◦ visual 26 Figure 16: Matching results using different parameters. Original recall fixations (in green) are matched to fixations during exploration (in yellow) and matched fixations are shown in fuchsia. Matching is stable across various parameters. The rightmost column shows cases of undesired deformation for extreme parameter settings. More fixations are considered for matching with a larger wp while the deformation is more rigid with a larger wd. Figure 17: The deformation model performs well on different examples. Parameters wp is set to 2◦ and wd to 10◦. Each example shows the fixations during exploration in yellow and recall fixations in green. Each pair overlaid with the stimulus is shown at the top-left corner. Matched fixations are shown in fuchsia. angle. This means we expect that matching fixations are usually not separated by more than twice this amount. To limit the deformation of the mapping we have tried values of wd ∈ [4◦, 16◦]. Based on experimentation we have settled for wp = 2◦ and wd = 10◦. Figure 17 shows the results of three examples using this parameter setting. In order to understand if deformation was really necessary to achieve the results we have presented, we perform the filtering based on the original, undeformed recall sequences. Similar to Figure 8, Figure 18 plots the percentage of remaining fixations as a function of the witness radius  varying from 1◦ to 10◦ visual angle, as well as the relative changes of importance of fixations. Comparing the results we notice that without applying deformation, much less fixations remain after filtering. For  = 1◦ 27 Figure 18: Results of filtering relative to witness radius  ∈ [1◦, 10◦] based on original recall fixations are shown in blue. No deformation mapping is applied. Results of filtering based on deformed recall fixations are depicted in orange for comparison. (a) Relative number of fixations retained after filtering. Each observer's data is scattered in the graph and the light color marks the center 50% region of the distribution. Much less fixations are retained when  is small. (b) Percentage of all stimuli maps whose most salient spots remain within 4◦ after filtering. More images contain changes of importance without applying deformation. Small  in particular highlights the effects of deformation mapping. nearly 20% of fixations during exploration remain after applying deformation, while only 5% have matched fixations in the original recall. Similar effects can be observed on the changes of the most important visual element: filtering based on the deformed recall sequences lead to significantly more of the images being unaffected. We spec- ulate that without deformation several important elements remain unmatched, causing significant but unwarranted changes to the map. 6.4. Visual attention and visual working memory Visual attention and visual working memory are entangled: attention controls en- coding in working memory by selecting what is the relevant information, and visual working memory biases the focus of attention (Downing, 2000; Schmidt et al., 2002; Olivers et al., 2006; Hollingworth et al., 2013). Several studies suggest that visual attention and visual working memory share the same representations (Farah, 1985; Harrison and Tong, 2009; Hollingworth and Hwang, 2013; Albers et al., 2013), and it has been debated whether they should be regarded as one cognitive function since they also share some of the same mechanisms (see Olivers (2008) for a review). 28 Our results suggest that the prioritization in visual attention differs from the priori- tization in working memory: the spatial importance maps are affected by filtering, and filtering is based on a functional connection to the spatial memory of visual elements. While more experiments are needed to understand this difference, we hope the exper- imental setup based on immediate recall of previously presented stimuli provides an additional useful tool. 7. Conclusions & Outlook The 'looking-at-nothing' phenomenon is an interesting way to test the awareness of a feature in an image previously fixated during exploration. This gives rise to a new way of measuring the importance of visual elements as assigned by a human observer. We have generated a data set based on eye tracking experiment that allows studying this idea and other researchers can use it as a basis for additional analysis. Overall, this importance model appears to conform to the idea of decomposing the image into relevant pieces - low-level features are only important for areas to be visited during exploration. We argue that classical saliency maps overemphasize the role of fixations. And concluding that objects which drew a lot attention based on fixations are indeed recognized by the observer may be wrong. The main challenge in this approach is the large amount of distortion, noise, and artifacts in the recall data. There seems to be two ways to address this problem: better processing of the data hopefully resulting in better matching; and better data collection. It might be possible to exploit the raw data to better match the recall to explo- ration sequence, as apparently not just the fixations correspond but also partial scan paths (Johansson et al., 2006). The matching might be improved using machine learn- ing techniques if we had ground truth data. It remains unclear how to generate such data. As mentioned, we have tried to get this information from the observers, but this attempt has been unsuccessful. We believe the experimental setup leaves little room for improvement in order to get cleaner data. The main problem is to motivate participants to equalize recall and eye movements without making the experiment significantly more complex. We made 29 one observation that would be worth exploiting: It appears that the recall performance improves in repeated trials. Similar effects have been observed in other recall experi- ments Mantyla and Holm (2006); Kaspar and Koenig (2011). A good protocol for such an experiment would still have to be designed. The setup clearly exploits short term memory. The results for long-timer recall, specifically if locations of fixations still encode the locations of features in images, are inconsistent (Martarelli and Mast, 2013; Laeng et al., 2014; Wantz et al., 2016). We expect the amount of noise in the data to increase and, therefore, finding the matching between recall and exploration to be even more difficult. Nonetheless, a setup similar to ours, yet with considerably delayed recall might be worthwhile. Such an experiment would be related to image memorability (Isola et al., 2014; Khosla et al., 2015; Bylin- skii et al., 2015), which explores the memorability of visual objects, and the results would likely become even more dependent on personal factors. The most successful techniques for predicting fixations to date are based on neural networks (Huang et al., 2015; Kruthiventi et al., 2015). We considered using similar approaches to predict the filtered heat maps. Better prediction would indicate that our results are more consistent. However, given the amount of noise in the data we use for filtering and the high-level effects we found in our analysis, we are unsure about the meaning of this analysis. We still think it would be generally fruitful to use filtered heat maps for learning important features that are closer to what humans really encode. Acknowledgments We would like to thank Marianne Maertens for valuable advice. Furthermore, we thank David Lindlbauer for help performing the experiment. This work has been par- tially supported by the ERC through grant ERC-2010-StG 259550 (XSHAPE). Appendix A. Computing the deformation mapping Let pi ∈ R2 be the positions of the fixations in the exploration sequence and rj ∈ R2 the positions of fixations in the recall sequence (in a common coordinate system). 30 We wish to compute a deformation D : R2 (cid:55)→ R2 that is applied to the recall locations rj with the aim to align the data with the fixations positions during exploration. We need two ingredients for computing the deformation: partial matching and de- formation mapping. For the deformation, we suggest Moving Least Squares (MLS) (Levin, 1998). Here, we use this framework applied to rigid transformations, i.e. local rigid transformations are fitted using weighted least squares. This approach has become popular in geometric modeling where it is usually derived as minimizing the deviation of the mapping from being locally isometric (Schaefer et al., 2006; Sorkine and Alexa, 2007; Chao et al., 2010). We model the deformation D as a rigid transformation that varies smoothly over space: D(x) = Rxx + tx. (A.1) The subscript x indicates that rotation and translation vary (smoothly) with the location x in the plane. They are computed by solving a weighted special orthogonal Procrustes problem (Gower and Dijksterhuis, 2004), where the weights depend on the distance of the points to x. Assume the desired position for rj is the position qj, then Rx, tx are computed by solving (cid:88) j arg min TRx=I,tx Rx θ((cid:107)x − rj(cid:107))(cid:107)Rxrj + tx − qj(cid:107)2 2. (A.2) Here, the weight function θ should be smoothly decaying with increasing distance. We use the common choice θd(x) = e − x2 w2 d , (A.3) which gives us control over the amount of deformation in the mapping with the parame- ter wd. The minimization can be solved directly using the singular value decomposition (SVD), see Sorkine-Hornung and Rabinovich (2016) for an accessible derivation. Note that for computing the mapping we simply assumed the desired positions qj were given. We compute them as the distance weighted centroid of exploration fixations: (cid:80) (cid:80) i θp((cid:107)D(rj) − pi(cid:107))pi i θp((cid:107)D(rj) − pi(cid:107)) qj = , (A.4) 31 Figure A.19: Iterations of the deformation. The initial state shows two fixation sets of viewing (orange) and recall (green). Radius of each circle correlates to the fixation duration. The set of fuchsia circles are the deformed recall fixations in each iteration. Note that the matching converges quickly after a few iterations. where θp(d) quickly decreases such that points further away are receiving relatively insignificant contribution. Note that we are considering the distances of the fixations pi to the deformed lo- cations of the fixations in the recall sequence. This means that setting the target loca- tions depends on the deformation mapping and computing the deformation mapping depends on the target locations. Consequently, we alternate the two steps as shown in Algorithm 1. We start this process with D being the identity. Then we compute the desired positions qj as explained above. The procedure converges after very few iterations (see Figure A.19). Note that the deformation D needs to be evaluated only in the location rj. This means for the next step we only need to compute Rx and tx for x = rj. References References Albers, A.M., Kok, P., Toni, I., Dijkerman, H.C., de Lange, F.P., 2013. Shared repre- sentations for working memory and mental imagery in early visual cortex. Current Biology 23, 1427 – 1431. doi:10.1016/j.cub.2013.05.065. Badcock, D.R., Hess, R.F., Dobbins, K., 1996. Localization of element clusters: Mul- tiple cues. Vision Research 36, 1467 – 1472. doi:10.1016/0042-6989(95) 00205-7. Bayliss, A.P., Murphy, E., Naughtin, C.K., Kritikos, A., Schilbach, L., Becker, S.I., 2013. "Gaze leading": Initiating simulated joint attention influences eye movements 32 Algorithm 1: Deformation mapping Data: Fixations locations in exploration p1, . . . , pn and recall r1, . . . , rm, convergence criteria λ 1, . . . , r(cid:48) Result: Locations of mapped recall fixations r(cid:48) 1 ∀(i, j) ∈ [1, m]2 : wij ← exp(−(cid:107)ri − rj(cid:107)2/w2 d) 2 for j ∈ [1, m] do 3 m (cid:17) (cid:16)(cid:80)m (cid:17) j ← rj r(cid:48) ¯ri ←(cid:16)(cid:80)m 6 repeat 4 5 7 8 9 10 11 12 13 j=1 wijrj / j=1 wij Ri = (wi1(r1 − ¯ri), . . . , wim(rm − ¯ri)) i−pj(cid:107)2/w2 p)pi i−pj(cid:107)2/w2 p) qi ← σ ← 0 for i ∈ [1, m] do (cid:80)n for i ∈ [1, m] do (cid:80)n j=1 exp(−(cid:107)r(cid:48) j=1 exp(−(cid:107)r(cid:48) ¯q ←(cid:16)(cid:80)m (cid:16)(cid:80)m (cid:17) UΣVT ← SVD(cid:0)RiQT(cid:1) Q = (q1 − ¯q, . . . , qm − ¯q) j=1 wijqj / j=1 wij (cid:17) 15 14 s ← VUT(ri − ¯ri) + ¯q σ ← σ + (cid:107)r(cid:48) i ← s r(cid:48) 17 until σ < λ i − s(cid:107) 16 33 and choice behavior. Journal of Experimental Psychology: General 142, 76–92. doi:10.1037/a0029286. Besl, P.J., McKay, N.D., 1992. Method for registration of 3-d shapes, in: Robotics-DL tentative, International Society for Optics and Photonics. pp. 586–606. Bochynska, A., Laeng, B., 2015. Tracking down the path of memory: eye scanpaths facilitate retrieval of visuospatial information. Cognitive Processing 16, 159–163. doi:10.1007/s10339-015-0690-0. Borji, A., Itti, L., 2015. Cat2000: A large scale fixation dataset for boosting saliency research. arXiv preprint arXiv:1505.03581 . Bruce, N.D.B., Tsotsos, J.K., 2005. Saliency based on information maximization, in: Proceedings of the 18th International Conference on Neural Information Processing Systems, MIT Press, Cambridge, MA, USA. pp. 155–162. Bylinskii, Z., Isola, P., Bainbridge, C., Torralba, A., Oliva, A., 2015. Intrinsic and extrinsic effects on image memorability. Vision Research 116, Part B, 165–178. doi:10.1016/j.visres.2015.03.005. computational Models of Visual At- tention. Bylinskii, Z., Judd, T., Oliva, A., Torralba, A., Durand, F., 2016. What do different evaluation metrics tell us about saliency models? arXiv preprint arXiv:1604.03605 . Chao, I., Pinkall, U., Sanan, P., Schroder, P., 2010. A simple geometric model for elas- tic deformations. ACM Trans. Graph. 29, 38:1–38:6. doi:10.1145/1778765. 1778775. Corkum, V., Moore, C., 1998. The origins of joint visual attention in infants. Develop- mental psychology 34, 28–38. doi:10.1037/0012-1649.34.1.28. Coutrot, A., Guyader, N., 2014. How saliency, faces, and sound influence gaze in dynamic social scenes. Journal of Vision 14, 5. doi:10.1167/14.8.5. 34 De Graef, P., Christiaens, D., d'Ydewalle, G., 1990. Perceptual effects of scene con- text on object identification. Psychological Research 52, 317–329. doi:10.1007/ BF00868064. Downing, P.E., 2000. Interactions between visual working memory and selective at- tention. Psychological Science 11, 467–473. doi:10.1111/1467-9280.00290. pMID: 11202491. Einhauser, W., Spain, M., Perona, P., 2008. Objects predict fixations better than early saliency. Journal of Vision 8, 18. doi:10.1167/8.14.18. Farah, M.J., 1985. Psychophysical evidence for a shared representational medium for mental images and percepts. Journal of Experimental Psychology: General 114, 91. doi:10.1037/0096-3445.114.1.91. Ferreira, F., Apel, J., Henderson, J.M., 2008. Taking a new look at looking at nothing. Trends in Cognitive Sciences 12, 405 – 410. doi:10.1016/j.tics.2008.07. 007. Gallup, A.C., Hale, J.J., Sumpter, D.J.T., Garnier, S., Kacelnik, A., Krebs, J.R., Couzin, I.D., 2012. Visual attention and the acquisition of information in human crowds. Proceedings of the National Academy of Sciences 109, 7245–7250. doi:10.1073/ pnas.1116141109. Gower, J.C., Dijksterhuis, G.B., 2004. Procrustes problems. volume 30. Oxford Uni- versity Press on Demand. Harrison, S.A., Tong, F., 2009. Decoding reveals the contents of visual working mem- ory in early visual areas. Nature 458, 632. doi:10.1038/nature07832. Hollingworth, A., Hwang, S., 2013. The relationship between visual working memory and attention: retention of precise colour information in the absence of effects on perceptual selection. Philosophical Transactions of the Royal Society of London B: Biological Sciences 368. doi:10.1098/rstb.2013.0061. 35 Hollingworth, A., Matsukura, M., Luck, S.J., 2013. Visual working memory modulates rapid eye movements to simple onset targets. Psychological Science 24, 790–796. doi:10.1177/0956797612459767. pMID: 23508739. Holmqvist, K., Nystrom, M., Andersson, R., Dewhurst, R., Halszka, J., van de Weijer, J., 2011. Eye Tracking: A Comprehensive Guide to Methods and Measures. Oxford University Press. Huang, X., Shen, C., Boix, X., Zhao, Q., 2015. Salicon: Reducing the semantic gap in saliency prediction by adapting deep neural networks, in: Proceedings of the IEEE International Conference on Computer Vision, pp. 262–270. Hutton, S.B., Nolte, S., 2011. The effect of gaze cues on attention to print advertise- ments. Applied Cognitive Psychology 25, 887–892. doi:10.1002/acp.1763. Illian, J., Penttinen, A., Stoyan, H., Stoyan, D., 2008. Statistical analysis and modelling of spatial point patterns. volume 70. John Wiley & Sons. Isola, P., Xiao, J., Parikh, D., Torralba, A., Oliva, A., 2014. What makes a photograph IEEE transactions on pattern analysis and machine intelligence 36, memorable? 1469–1482. Itti, L., Koch, C., 2000. A saliency-based search mechanism for overt and covert shifts of visual attention. Vision research 40, 1489–1506. Jiang, M., Huang, S., Duan, J., Zhao, Q., 2015. Salicon: Saliency in context, in: Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, pp. 1072–1080. Johansson, R., Holsanova, J., Dewhurst, R., Holmqvist, K., 2012. Eye movements dur- ing scene recollection have a functional role, but they are not reinstatements of those produced during encoding. Journal of Experimental Psychology: Human Perception and Performance 38, 1289–1314. Johansson, R., Holsanova, J., Holmqvist, K., 2006. scriptions elicit similar eye movements during mental Pictures and spoken de- imagery, both in light 36 and in complete darkness. Cognitive Science 30, 1053–1079. doi:10.1207/ s15516709cog0000_86. Johansson, R., Johansson, M., 2014. Look here, eye movements play a functional role in memory retrieval. Psychological Science 25, 236–242. Judd, T., Durand, F., Torralba, A., 2012. A benchmark of computational models of saliency to predict human fixations. Technical Report MIT-CSAIL-TR-2012-001. MIT. Judd, T., Ehinger, K., Durand, F., Torralba, A., 2009. Learning to predict where humans look, in: Computer Vision, 2009 IEEE 12th international conference on, IEEE. pp. 2106–2113. Kaspar, K., Koenig, P., 2011. Viewing behavior and the impact of low-level image properties across repeated presentations of complex scenes. Journal of Vision 11, 26–26. Khosla, A., Raju, A.S., Torralba, A., Oliva, A., 2015. Understanding and predicting image memorability at a large scale, in: Proceedings of the IEEE International Con- ference on Computer Vision, pp. 2390–2398. Kienzle, W., Franz, M.O., Scholkopf, B., Wichmann, F.A., 2009. Center-surround patterns emerge as optimal predictors for human saccade targets. Journal of vision 9, 7–7. Kim, N.W., Bylinskii, Z., Borkin, M.A., Gajos, K.Z., Oliva, A., Durand, F., Pfister, H., 2017. Bubbleview: an alternative to eye-tracking for crowdsourcing image impor- tance. arXiv preprint arXiv:1702.05150 . Kim, N.W., Bylinskii, Z., Borkin, M.A., Oliva, A., Gajos, K.Z., Pfister, H., 2015. A crowdsourced alternative to eye-tracking for visualization understanding, in: Pro- ceedings of the 33rd Annual ACM Conference Extended Abstracts on Human Fac- tors in Computing Systems, ACM. pp. 1349–1354. 37 Koehler, K., Guo, F., Zhang, S., Eckstein, M.P., 2014. What do saliency models pre- dict? Journal of vision 14, 14–14. doi:10.1167/14.3.14. Koivisto, M., Hyona, J., Revonsuo, A., 2004. The effects of eye movements, spatial attention, and stimulus features on inattentional blindness. Vision Research 44, 3211 – 3221. doi:10.1016/j.visres.2004.07.026. Krieger, G., Rentschler, I., Hauske, G., Schill, K., Zetzsche, C., 2000. Object and scene analysis by saccadic eye-movements: an investigation with higher-order statistics. Spatial Vision 13, 201–214. doi:10.1163/156856800741216. Kruthiventi, S.S., Ayush, K., Babu, R.V., 2015. Deepfix: A fully convolutional neural network for predicting human eye fixations. arXiv preprint arXiv:1510.02927 . Laeng, B., Bloem, I.M., DAscenzo, S., Tommasi, L., 2014. Scrutinizing visual images: The role of gaze in mental imagery and memory. Cognition 131, 263–283. Laeng, B., Teodorescu, D.S., 2002. Eye scanpaths during visual imagery reenact those of perception of the same visual scene. Cognitive Science 26, 207–231. doi:10. 1207/s15516709cog2602_3. Levin, D., 1998. The approximation power of moving least-squares. Math. Comput. 67, 1517–1531. doi:10.1090/S0025-5718-98-00974-0. Loftus, G., Mackworth, N., 1978. Cognitive determinants of fixation location dur- ing picture viewing. Journal of Experimental Psychology: Human Perception and Performance 4, 565–572. doi:10.1037/0096-1523.4.4.565. Mantyla, T., Holm, L., 2006. Gaze control and recollective experience in face recogni- tion. Visual Cognition 14, 365–386. Martarelli, C.S., Chiquet, S., Laeng, B., Mast, F.W., 2017. Using space to repre- sent categories: insights from gaze position. Psychological Research 81, 721–729. doi:10.1007/s00426-016-0781-2. Martarelli, C.S., Mast, F.W., 2013. Eye movements during long-term pictorial recall. Psychological Research 77, 303–309. doi:10.1007/s00426-012-0439-7. 38 Matin, E., 1974. Saccadic suppression: a review and an analysis. Psychological bulletin 81, 899. Memmert, D., 2006. The effects of eye movements, age, and expertise on inatten- tional blindness. Consciousness and Cognition 15, 620 – 627. doi:10.1016/j. concog.2006.01.001. Moore, C.S., 1903. Control of the memory image. The Psychological Review: Mono- graph Supplements . Olivers, C.N., 2008. Interactions between visual working memory and visual attention. Frontiers in Bioscience 13, 1182–1191. doi:10.2741/2754. Olivers, C.N., Meijer, F., Theeuwes, J., 2006. Feature-based memory-driven at- tentional capture: visual working memory content affects visual attention. Jour- nal of Experimental Psychology: Human Perception and Performance 32, 1243. doi:10.1037/0096-1523.32.5.1243. Oztireli, A.C., Gross, M., 2012. Analysis and synthesis of point distributions based on pair correlation. ACM Trans. Graph. 31, 170:1–170:10. doi:10.1145/2366145. 2366189. Pannasch, S., Helmert, J.R., Roth, K., Herbold, A.K., Walter, H., 2008. Visual fixation durations and saccade amplitudes: Shifting relationship in a variety of conditions. Journal of Eye Movement Research 2. Parkhurst, D., Law, K., Niebur, E., 2002. Modeling the role of salience in the allo- cation of overt visual attention. Vision Research 42, 107 – 123. doi:10.1016/ S0042-6989(01)00250-4. Pathman, T., Ghetti, S., 2015. Eye movements provide an index of veridical mem- ory for temporal order. PLOS ONE 10, 1–17. doi:10.1371/journal.pone. 0125648. Ramanathan, S., Katti, H., Sebe, N., Kankanhalli, M., Chua, T.S., 2010. An eye fixation database for saliency detection in images, in: Daniilidis, K., Maragos, 39 P., Paragios, N. (Eds.), Computer Vision – ECCV 2010: 11th European Con- ference on Computer Vision, Heraklion, Crete, Greece, September 5-11, 2010, Proceedings, Part IV, Springer Berlin Heidelberg, Berlin, Heidelberg. pp. 30–43. doi:10.1007/978-3-642-15561-1_3. Recasens, A., Khosla, A., Vondrick, C., Torralba, A., 2015. Where are they looking?, in: Cortes, C., Lawrence, N.D., Lee, D.D., Sugiyama, M., Garnett, R. (Eds.), Ad- vances in Neural Information Processing Systems 28. Curran Associates, Inc., pp. 199–207. Ricciardelli, P., Bricolo, E., Aglioti, S.M., Chelazzi, L., 2002. My eyes want to look where your eyes are looking: Exploring the tendency to imitate another individual's gaze. NeuroReport 13, 2259–2264. Schaefer, S., McPhail, T., Warren, J., 2006. Image deformation using moving least squares, in: ACM transactions on graphics (TOG), ACM. pp. 533–540. Schmidt, B.K., Vogel, E.K., Woodman, G.F., Luck, S.J., 2002. Voluntary and automatic attentional control of visual working memory. Perception & Psychophysics 64, 754– 763. doi:10.3758/BF03194742. Scholz, A., von Helversen, B., Rieskamp, J., 2015. Eye movements reveal memory processes during similarity- and rule-based decision making. Cognition 136, 228 – 246. doi:10.1016/j.cognition.2014.11.019. Scholz, A., Mehlhorn, K., Krems, J.F., 2016. Listen up, eye movements play a role in verbal memory retrieval. Psychological Research 80, 149–158. doi:10.1007/ s00426-014-0639-4. Schomaker, J., Walper, D., Wittmann, B.C., Einhauser, W., 2017. Attention in natural scenes: Affective-motivational factors guide gaze independently of visual salience. Vision Research 133, 161 – 175. doi:10.1016/j.visres.2017.02.003. Simons, D.J., 2000. Attentional capture and inattentional blindness. Trends in Cogni- tive Sciences 4, 147 – 155. doi:10.1016/S1364-6613(00)01455-8. 40 Sorkine, O., Alexa, M., 2007. As-rigid-as-possible surface modeling, in: Proceed- ings of the Fifth Eurographics Symposium on Geometry Processing, Eurographics Association, Aire-la-Ville, Switzerland, Switzerland. pp. 109–116. Sorkine-Hornung, O., Rabinovich, M., 2016. Least-squares rigid motion using svd. Technical note. Tatler, B.W., 2007. The central fixation bias in scene viewing: Selecting an opti- mal viewing position independently of motor biases and image feature distributions. Journal of Vision 7, 4. doi:10.1167/7.14.4. Tatler, B.W., Hayhoe, M.M., Land, M.F., Ballard, D.H., 2011. Eye guidance in natural vision: Reinterpreting salience. Journal of vision 11, 5–5. doi:10.1167/11.5.5. Tavakoli, H.R., Ahmed, F., Borji, A., Laaksonen, J., 2017. Saliency revisited: Analysis of mouse movements versus fixations. arXiv preprint arXiv:1705.10546 . Trukenbrod, H.A., Barthelm´e, S., Wichmann, F.A., Engbert, R., 2017. Rigorous spatial statistics for gaze patterns in scene viewing: Effects of repeated viewing. arXiv preprint arXiv:1704.01761 . Van Der Linde, I., Rajashekar, U., Bovik, A., Cormack, L., 2009. Doves: a database of visual eye movements. Spatial Vision 22, 161–177. doi:10.1163/ 156856809787465636. de Vito, S., Buonocore, A., Bonnefon, J.F., Della Sala, S., 2014. Eye movements disrupt spatial but not visual mental imagery. Cognitive processing 15, 543–549. Vo, M.L.H., Aizenman, A.M., Wolfe, J.M., 2016. You think you know where you looked? you better look again. Journal of experimental psychology: human percep- tion and performance 42, 1477. Wantz, A.L., Martarelli, C.S., Mast, F.W., 2016. When looking back to noth- ing goes back to nothing. Cognitive Processing 17, 105–114. doi:10.1007/ s10339-015-0741-6. 41 Xu, J., Jiang, M., Wang, S., Kankanhalli, M.S., Zhao, Q., 2014. Predicting human gaze beyond pixels. Journal of vision 14, 28–28. 42
1512.08309
2
1512
2016-08-16T19:21:32
Identifying Seizure Onset Zone from the Causal Connectivity Inferred Using Directed Information
[ "q-bio.NC", "cs.IT", "cs.IT", "stat.ME" ]
In this paper, we developed a model-based and a data-driven estimator for directed information (DI) to infer the causal connectivity graph between electrocorticographic (ECoG) signals recorded from brain and to identify the seizure onset zone (SOZ) in epileptic patients. Directed information, an information theoretic quantity, is a general metric to infer causal connectivity between time-series and is not restricted to a particular class of models unlike the popular metrics based on Granger causality or transfer entropy. The proposed estimators are shown to be almost surely convergent. Causal connectivity between ECoG electrodes in five epileptic patients is inferred using the proposed DI estimators, after validating their performance on simulated data. We then proposed a model-based and a data-driven SOZ identification algorithm to identify SOZ from the causal connectivity inferred using model-based and data-driven DI estimators respectively. The data-driven SOZ identification outperforms the model-based SOZ identification algorithm when benchmarked against visual analysis by neurologist, the current clinical gold standard. The causal connectivity analysis presented here is the first step towards developing novel non-surgical treatments for epilepsy.
q-bio.NC
q-bio
Identifying Seizure Onset Zone from the Causal Connectivity Inferred Using Directed Information Rakesh Malladi, Student Member, IEEE, Giridhar Kalamangalam, Nitin Tandon, and Behnaam Aazhang, Fellow, IEEE 1 6 1 0 2 g u A 6 1 ] . C N o i b - q [ 2 v 9 0 3 8 0 . 2 1 5 1 : v i X r a Abstract -- In this paper, we developed a model-based and a data-driven estimator for directed information (DI) to infer the causal connectivity graph between electrocorticographic (ECoG) signals recorded from brain and to identify the seizure onset zone (SOZ) in epileptic patients. Directed information, an information theoretic quantity, is a general metric to infer causal connectivity between time-series and is not restricted to a particular class of models unlike the popular metrics based on Granger causality or transfer entropy. The proposed estimators are shown to be almost surely convergent. Causal connectivity between ECoG electrodes in five epileptic patients is inferred using the proposed DI estimators, after validating their performance on simulated data. We then proposed a model-based and a data-driven SOZ identification algorithm to identify SOZ from the causal connec- tivity inferred using model-based and data-driven DI estimators respectively. The data-driven SOZ identification outperforms the model-based SOZ identification algorithm when benchmarked against visual analysis by neurologist, the current clinical gold standard. The causal connectivity analysis presented here is the first step towards developing novel non-surgical treatments for epilepsy. Index Terms -- Epilepsy, directed information, seizure onset zone, ECoG, causal connectivity. I. INTRODUTION Epilepsy is a common neurological disease affecting nearly 1% of the world's population. Epilepsy is characterized by unprovoked seizures, which are periods of hypersynchronous activity in the brain. The current treatment options include medication, resective surgery and more recently, electrical stimulation approaches like vagus nerve and responsive neuro stimulation. However, medication is not able to stop seizures in about one-third of the patients. The efficacy of the other current neuro-modulation approaches is variable and almost never results in a cure [3], [4]. The current approaches lack specificity and suffer from negative side effects ( [5] and ref- erences therein). Selective modulation of the epileptic circuits in the brain via electrical stimulation [6], optogenetics and designer receptive technologies [5] represent possible options Copyright (c) 2016 IEEE. Personal use of this material is permitted. However, permission to use this material for any other purposes must be obtained from the IEEE by sending a request to [email protected]. This work is funded in part by grant 1406447 from National Science Foundation and Texas Instruments. A portion of this work was presented at Society of Neuroscience (SfN) 2014 [1] and International Conference on Acoustics, Speech and Signal Processing (ICASSP) 2015 [2]. Rakesh Malladi and Behnaam Aazhang are with the Department of Elec- trical and Computer Engineering, Rice University, Houston, TX, 77005 USA. Giridhar Kalamangalam and Nitin Tandon are with Department of Neu- rology and Department of Neurosurgery, respectively, at University of Texas Health Center, Houston, TX, 77005. E-mail: {rm17, aaz}@rice.edu, {Giridhar.P.Kalamangalam, Nitin.Tandon}@uth.tmc.edu. for better treatments for this disabling disease. A crucial initial step in this endeavor is understanding how seizures originate and spread within the brain. Effective or causal connectivity [7] quantifies how the activity spreads between different brain regions and can be used to characterize epileptic networks. In addition, causal connectivity can also be used to identify seizure onset zone (SOZ) (brain regions initiating seizures [8]) and has been shown to predict the efficacy of resective surgery [9], [10]. The main objective of this paper is to develop estimators of directed information (DI), derive causal connectivity between brain regions and identify the SOZ using electrocorticographic (ECoG) data in patients with epilepsy. Estimating causal connectivity from electrophysiological recordings of brain has been the focus of many papers. A good summary is provided in [9]. The causality referred to in this paper is in the Wiener-Granger causal sense [11]. Metrics based on Granger causality (GC) [12], [13] and information theory like transfer entropy [14] are commonly used to es- timate causal connectivity between continuous-valued ECoG recordings. However, these techniques are well-suited only for a specific model and subset of the recorded signals. For instance, GC-based measures are applicable only for data from multivariate autoregressive (MVAR) processes. ECoG data are often modeled using linear MVAR model [9], [13], even though associations between seizure loci detected by ECoG recordings are likely nonlinear [15], [16]. We propose to develop a causal metric that would be applicable to diverse models and different types of electro- physiological recordings of brain. Directed information, used to infer causal connections between spike trains in [17] -- [19], can indeed be further developed into a general technique to estimate causal connectivity. The definition of DI is based on the underlying probability distribution and no assumptions are imposed on the underlying distributions. Directed information was developed for discrete-valued time-series in [20] -- [22] and nonparametrically estimated in [23]. DI quantifies the amount of causal information about one time-series that is explained by the other time-series [17]. Modified time-lagged directed information is proposed in [24], [25] to reduce the computational complexity of estimating directed information. DI is also used in many other applications [26] -- [28]. The definition of DI is broadened to the class of continuous-valued processes like ECoG signals in this paper [1], [24], [25], [29]. If the data is assumed to be from a MVAR process with Gaussian white noise, DI is equivalent to Granger causality [29] and if the data satisfies the general Markov condition, DI is very closely related to transfer entropy [24], [25]. The main 2 advantage of DI over other existing techniques is that DI is applicable to a large class of electrophysiological recordings from brain, including spike trains, EEG and ECoG, and is not restricted to a particular class of models. We developed an almost surely convergent model-based and data-driven estimators of DI in this paper, inspired by prior work [17], [30]. The performance of the proposed DI estimators was validated on linear and nonlinear simulated models and compared with the Granger causality metric [12], [31]. The statistical significance of the causal connection inferred using DI and GC estimates was demonstrated using an adaptation [32] of stationary bootstrap [33]. We then applied both model-based and data-driven DI estimators to infer causal connectivity graphs between ECoG channels from twelve seizures in five patients with epilepsy. The DI metric with model-based and data-driven estimators proposed in this paper allows us the flexibility to simultane- ously use both these estimators and identify which one leads to more reliable causal connectivity graphs from real ECoG data. This would also allows us to examine the appropriateness of imposing linear MVAR assumption on ECoG data. We used the model-based DI estimator with MVAR model assumption to detect the linear causal interactions and the data-driven DI estimator to detect both linear and nonlinear causal interactions between ECoG channels. We observed that nonlinear causal interactions between channels are stronger around the onset of a seizure, as widely believed [16]. We then proposed a model-based and a data-driven SOZ identification algorithm to identify SOZ from the causal con- nectivity graphs inferred using model-based and data-driven DI estimators respectively. The SOZ identified by model-based and data-driven algorithms are respectively the isolated nodes and strong sources in the corresponding causal connectivity graphs. Despite the numerous SOZ identification algorithms available [9], [10], [34] -- [36], the current clinical gold standard is still the visual analysis of ECoG data by the neurologist. We therefore compared the performance of both model-based and data-driven SOZ identification algorithms with visual analysis by the neurologist. We find that the data-driven approach outperforms the model-based approach and also leads to more interpretable results. The methodology proposed here should be extended to analyze the whole ECoG record to better understand the evolution of seizure mechanisms over time. The main algorithmic contributions of this paper are • Developing an almost surely convergent model-based and data-driven DI estimator, described in sections III and IV. • Developing a MVAR model-based and a data-driven SOZ identification algorithm, described in section VI. II. DIRECTED INFORMATION Consider the N samples recorded at a sampling frequency Fs from each ECoG electrode implanted in an epileptic patient. Without let us focus on two channels X and Y. The N samples recorded from the two channels are denoted by XN = (x1, x2,··· , xN )T and YN = (y1, y2,··· , yN )T. Also let W denote the matrix of samples recorded from a group of channels excluding X and Y. For loss of generality, notational simplicity, the elements corresponding to the non- positive subscripts are treated as empty sets and the subscripts are not shown when equal to 1. The directed information, I(cid:0)XN → YN(cid:1), from N samples of continuous-valued random process X to those of Y is defined as I(cid:0)XN → YN(cid:1) = h(cid:0)YN(cid:1) − h(cid:0)YN(cid:107)XN(cid:1) , where h(cid:0)YN(cid:1) is the differential entropy of the N-dimensional continuous random vector YN [30] and h(cid:0)YN(cid:107)XN(cid:1) is the (1) causally conditioned differential entropy of YN causally con- ditioned on XN . The causally conditioned differential entropy is defined as h(cid:0)YN(cid:107)XN(cid:1) = N(cid:80) h(cid:0)ynYn−1, Xn(cid:1) . (2) n=1 of bits of uncertainty in one process that is causally explained The definitions of DI and causally conditioned differential en- tropy in (1) and (2) are obtained by broadening the definitions of the same quantities from discrete-time, discrete-valued ran- dom processes [21], [22] to discrete-time, continuous-valued processes [2]. One of the main differences between discrete- valued and continuous-valued random processes is that the entropy of a discrete-valued process is always non-negative, whereas the differential entropy of a continuous-valued pro- cess can be negative [30]. However, DI is always non-negative since conditioning cannot increase differential entropy [30], i.e., I(cid:0)XN → YN(cid:1) ≥ 0. DI can be interpreted as the number away by the other process. If I(cid:0)XN → YN(cid:1) = 0, then there metric in general, i.e., I(cid:0)XN → YN(cid:1) (cid:54)= I(cid:0)YN → XN(cid:1). mutual information [30], I(cid:0)yn; XnYn−1(cid:1), as I(cid:0)XN → YN(cid:1) = (cid:8)h(cid:0)ynYn−1(cid:1) − h(cid:0)ynYn−1, Xn(cid:1)(cid:9) I(cid:0)yn; XnYn−1(cid:1) . N I(cid:0)XN→ YN(cid:1) N h(cid:0)YN(cid:1)−limN→∞ 1 N(cid:80) N(cid:80) I(cid:0)X→ Y(cid:1) =limN→∞ 1 Note that DI can also be expressed in terms of conditional is no causal influence from X to Y. The DI is not a symmetric Now, the DI between the time-series X and Y is defined as N h(cid:0)YN(cid:107)XN(cid:1) (3) n=1 n=1 = = h(cid:0)Y(cid:1)−h(cid:0)Y(cid:107)X(cid:1), =limN→∞ 1 (4) provided the limits exist. h (Y) and h (Y(cid:107)X) are respectively the differential entropy of Y and the causally conditioned differential entropy of Y given X. The DI from Y to X is also similarly defined. Furthermore, the DI defined earlier is easily extended to define directed information from X to Y, causally conditioned on W. Note that W comprises the samples recorded from a group of ECoG channels excluding X and Y. The causally conditioned DI, I (X → Y(cid:107)W), is defined as I (X → Y(cid:107) W) = limN→∞ 1 N I(cid:0)XN → YN(cid:107)WN(cid:1) (cid:8)h(cid:0)YN(cid:107)WN(cid:1)−h(cid:0)YN(cid:107)XN,WN(cid:1)(cid:9) , = limN→∞ 1 N = h (Y(cid:107)W) − h (Y(cid:107)W, X), (5) where h(cid:0)YN(cid:107)XN, WN(cid:1) = N(cid:80) h(cid:0)ynYn−1, Xn, Wn(cid:1) is the n=1 differential entropy of YN causally conditioned on XN and WN , h (Y(cid:107)W) is the causal conditioned differential entropy of Y given W and h (Y(cid:107)W, X) is the causally conditioned differential entropy of Y given the causal past of W and X. We use the directed information defined here to learn the causal connectivity graph between all ECoG channels and identify the SOZ of epileptic patients. III. UNIVERSAL ESTIMATOR FOR DIRECTED INFORMATION A universal estimator for directed information between channels X and Y, I (X → Y), and the causally conditioned DI, I (X → Y(cid:107)W) is developed in this section. The proposed estimator is universal and is shown to be almost surely convergent assuming that the causal conditional likelihood (CCL) is known. If CCL is not known and is estimated, then the convergence of the proposed DI estimator is dependent on the CCL estimator. The ideas used in developing the proposed DI estimator are inspired by prior work [17], [30]. Without loss of generality, we will first focus on estimating the pairwise DI, I (X → Y). We will then outline the procedure to extend this pairwise DI estimator to estimate the causally conditioned DI, I (X → Y(cid:107)W). The inputs to the proposed pairwise DI estimator are the observed N samples of time-series X and Y. The main idea is to develop an almost surely convergent estimator for the entropies in (4) and the difference between the two entropy estimates is an almost surely convergent estimate for I (X → Y). Let us first focus on the causally conditioned differential entropy estimator, h (Y(cid:107)X). Assumption 1 - The random processes X and Y are assumed to be stationary, ergodic and Markovian in the observed time- window. These are reasonable assumptions to model ECoG data. First, an implicit assumption in the problem of estimating the causal connectivity from a ECoG data segment is that the causal connectivity does not vary in this segment, which is mathematically captured by stationarity. The entire ECoG data record is usually not stationary and stationary segments are identified using either sliding windows [37] or change- point detection algorithms [38]. We used the sliding window approach in this paper. A crucial parameter in this process is the length of the window in which data is assumed to be stationary. It is also important to realize that we need a mini- mum amount of data points to reliably estimate any unknown parameters involved. It is recommended that the number of data points should be much larger (as a thumb rule, at least an order of magnitude larger) than the number of parameters to be estimated [37]. Directed information is then estimated in each stationary segment using the algorithm proposed in this section. Directed information for the entire time-series is the sum of the DI estimates from each stationary segment and is interpreted as the total amount of uncertainty in one time- series in the entire recording window that is explained by the other time-series. Second, ergodicity is required to ensure that the estimates from long-enough recording windows converge to the true value. Finally, the Markovian assumption captures the dependence of the current activity on the past activity 3 at different electrodes. Let the current sample of the time- series Y depend on the past Jyy and past Kyx samples of the time-series Y and X respectively. Note that (Jyy, Kyx) are unknown and should be estimated from data. The explicit model of the dependence is captured by the causal likelihood of yn conditioned on the past activity at electrodes X and Y. This CCL is denoted by P(cid:0)ynYn−1 (cid:1) and , Xn can be estimated using either a model-based or a data-driven approach. Let us assume for now that CCL is known. n−Kyx+1 n−Jyy , Xl (cid:16) and Xl ylYl−1 l−Jyy (cid:17) ∈ R. Assumption 2 - Let us also assume that differential entropy of the first sample, y1, of time-series Y exists and that for some time-index l ∈ [1, N ], the conditional differential entropy of yl, conditioned on Yl−1 l−Kyx+1 also exists, i.e., l−Jyy h (y1) , h l−Kyx+1 Lemma 3.1. Let Assumptions 1 and 2 hold. Then h (Y) , h (Y(cid:107)X) , I (X → Y) exists and are in R. Proof. Stationarity and the property that conditioning cannot increase the differential entropy are the main ideas in the proof, which is in the Appendix A. Lemma 3.2. Let Assumptions 1 and 2 hold. Then for some time-index l N h(cid:0)YN(cid:107)XN(cid:1) =E(cid:104)−log P (cid:16) ylYl−1 l−Jyy l−(Kyx−1) (cid:17)(cid:105) . (6) , Xl 1 Proof. The proof uses the definition of causally conditioned differential entropy (2), the Markovian and the stationarity assumptions. The proof is in the Appendix A. Theorem 3.1. Let Assumptions 1 and 2 hold. Then the almost surely convergent causally conditioned differential entropy estimator is h (Y(cid:107)X) = 1 (cid:110)−log P (cid:16) (cid:17)(cid:111) N(cid:80) , Xn . (7) ynYn−1 n−Jyy n−(Kyx−1) N n=1 Proof. The proof is based on two observations: the first is that the right-hand side of (6) does not depend on N and therefore it is easy to compute its limit as N → ∞. The second observation is that the strong law of large numbers (SLLN) for Markov chains [39] can be applied to estimate the expectation on the right-hand side of (6). The detailed proof is in the Appendix A. An almost surely convergent estimator for h (Y) can be easily derived using Theorem 3.1, simply by modeling the dependence of the current samples of Y on its own J(cid:48) yy past samples. This is equivalent to setting Kyx = 0. The difference between the two estimators, h (Y) and h (Y(cid:107)X), is the almost surely convergent estimator for DI from X to Y, I (X → Y). This is stated in Theorem 3.2. Theorem 3.2. Let Assumptions 1 and 2 hold. The universal estimator for DI from time-series X to Y is I (X → Y) = h (Y) − h (Y(cid:107)X) a.s.−−→ I (X → Y) . (8) Proof. We have from Theorem 3.1 I (X → Y) = h (Y) − h (Y(cid:107)X) a.s.−−→ h (Y) − h (Y(cid:107)X) = I (X → Y) . The DI estimator in Theorem 3.2 can be easily extended to estimate the causally conditioned directed information, 4 (cid:16) I (X → Y(cid:107)W). First, h (Y(cid:107)W) is estimated using Theo- rem 3.1. We now need to estimate h (Y(cid:107)W, X). Let Jyy, Kyw and Kyx respectively denote the number of past sam- ples of Y, W and X that influence the current sample of Y. Let us also assume the causal conditional likelihood is known. A model- P based and a data-driven approach to estimate this CCL is described in the subsequent section. Then Theorem. 3.1 can be easily extended to show that ynYn−1 n−Jyy n−Kyw+1, Xn n−Kyx+1 , Wn (cid:17) h (Y(cid:107)W, X) N(cid:80) (cid:110)−log P (cid:16) = 1 N n=1 ynYn−1 n−Jyy ,Wn n−Kyw+1,Xn n−Kyx+1 (cid:17)(cid:111) (9) is important there is no causal information X contains about Y. It is an almost surely convergent estimate of h (Y(cid:107)W, X). From (5), I (X → Y(cid:107)W) is the difference between the estimates, h (Y(cid:107)W) and h (Y(cid:107)W, X). It to note that the as the number of channels included in W increases, computational complexity of the estimator also increases. The DI estimate, I (X → Y), can be interpreted as the amount of causal is, however, important to note that I (X → Y) is estimated from N samples and is an estimate of the true value of DI from X to Y. The statistical significance of the causal connection from X to Y inferred from I (X → Y) is calculated using an adaptation [32] of stationary bootstrap [33]. B stationary bootstrap samples of X, denoted by X(b), are generated using the algorithm described in [32] for b = 1, 2,··· , B. The DI from bth stationary bootstrap sample X(b) to Y, denoted by I (X(b) → Y), is estimated using the proposed DI estimator. Note that influence from any of these bootstrap samples to Y by construction. Therefore the B samples, I (X(b) → Y), for b = 1, 2,··· , B are from the null hypothesis of no causal influence. The statistical significance is determined by the P-value [40]. P-value is the probability that DI estimate greater than or equal to I (X → Y) can be observed under the null hypothesis of no causality from X to Y and is computed from the empirical distribution of I (X(b) → Y) for b = 1, 2,··· , B. If the P-value is less than a predetermined significance level δ, the null hypothesis of no causal connection from X to Y is rejected. On the other hand, if the P-value is greater than δ, the null hypothesis cannot be rejected and the causal connection from X to Y is not statistically significant. Note that the empirical distribution of I (X(b) → Y) is concentrated around 0, since the DI between time-series that are not causally connected is zero. Therefore, when the actual DI estimate is large enough, the P-value will be less than δ and the statistical significance assessment is not required. However, statistical significance assessment is useful when the DI estimate is close to zero. The significance assessment described here is applied to the simulated examples in section V to identify the significant causal connections, particularly useful when the DI estimates are close to zero. The above discussion assumes CCL is known. The likelihood, however, must be estimated from data in practice. A model- based and a data-driven approach to estimate CCL is described in the following section. IV. ESTIMATING CAUSAL CONDITIONAL LIKELIHOOD Estimating DI from X to Y using the proposed DI estimator in section III requires estimating two CCLs, P(cid:0)ynYn−1(cid:1) and P(cid:0)ynYn−1, Xn(cid:1), while estimating DI from Y to X P(cid:0)ynYn−1, Wn(cid:1) and P(cid:0)ynYn−1, Wn, Xn(cid:1). Let us focus on estimating P(cid:0)ynYn−1, Xn(cid:1) for n = 1, 2,··· , N, which to extend this approach to estimate P(cid:0)ynYn−1, Wn, Xn(cid:1). is required to estimate h (Y(cid:107)X). We will then describe how causally conditioned on W requires estimating two CCLs, The CCLs are estimated using either model-based or data- driven techniques. The choice between model-based and data- driven approaches is determined by the application from which data is recorded. For instance, the time-series signals obtained from electrophysiological recordings of brain or from stock markets are commonly modeled using MVAR models with Gaussian white noise. In this case, the CCL is easily estimated from the MVAR model of the data. Usually the parameters of the model are unknown and several classical techniques to estimate the unknown parameters are described in [41]. On the other hand, using model-based approaches to estimate CCLs from data recorded from nonlinear systems or systems without a prescribed linear model is non-trivial. This is because estimating the CCLs using model-based approach requires essentially inverting the nonlinear generative model, which is not trivial. Data-driven approaches do not have this limitation and are therefore preferred for nonlinear time-series data. A good review of the various data-driven algorithms that estimate probability distribution from data is provided in [42], [43]. The model-based and the data-driven CCL algorithm used in this paper are described in the remainder of this section. A. Model-based CCL Estimation We will focus on estimating the CCL specifically for multivariate autoregressive process with Gaussian white noise in this paper. Let the time-series X and Y be sampled from such processes. Then, the samples of Y can be expressed as Jyy(cid:80) Kyx(cid:80) yn = αjyn−j + βkxn−k+1 +zn,n = 1, 2,··· ,N, (10) j=1 k=1 where zn is the additive white Gaussian noise with zero mean z. Here αj for j = 1, 2,··· , Jyy and βk for k = and variance σ2 1, 2,··· , Kyx are the parameters of the model and Jyy and Kyx are the model orders representing how many past samples of Y and of X respectively influence the current sample of Y. It is easy to observe from (10) that P(cid:0)ynYn−1,Xn(cid:1)∼N(cid:16)Jyy(cid:80) (cid:17) vector θ (Jyy, Kyx) =(cid:0)α1,··· , αJyy , β1,··· , βKyx, σ2 The two model orders, Jyy and Kyx, and the parameter (cid:1)T are βkxn−k+1, σ2 z Kyx(cid:80) αjyn−j + not known apriori and need to be estimated from the N observed samples of X and Y. The parameters and the model orders are estimated using a maximum likelihood (ML) estimator with minimum description length [44] penalty. ML estimator is known to be asymptotically consistent. Mini- mum description length is a model order selection procedure (11) k=1 j=1 . z to solutions of the following problem: (cid:1) = arg min with good consistency properties [44] and proportional (Jyy + Kyx). The optimal model orders (cid:0) Jyy, Kyx (cid:1) are the N logP(cid:0)YN(cid:107)XN;θ(cid:0)Jyy, Kyx (cid:1)(cid:1) (cid:0) Jyy, Kyx (cid:8)− 1 log N(cid:9), negative log-likelihood for a given(cid:0)Jyy, Kyx (cid:1) and is obtained N logP(cid:0)YN(cid:107)XN;θ (Jyy,Kyx)(cid:1) . (13) (12) where θ (Jyy, Kyx) is the value of θ which minimizes the by solving θ (Jyy, Kyx)=arg min + Jyy+Kyx (Jyy,Kyx) − 1 2N θ The ML estimation of θ for a given (Jyy, Kyx) in (13) is equivalent to the ML estimation of the parameters of a standard linear regression model [41], since the CCL is Gaussian distributed (11). The estimated parameters almost surely converge to the true parameter values [45] resulting in almost surely convergence of the proposed DI estimator. The desired CCL is obtained by substituting the solutions of (13), (12) in (11). The resultant CCL is then substituted in (7) to estimate h (Y(cid:107)X), which is further simplified to h (Y(cid:107)X) = 1 z is the estimate of the noise variance from (12), (13). 2 log(cid:0)2πeσ2 (cid:1), where σ2 Let us focus on estimating P(cid:0)ynYn−1, Wn, Xn(cid:1), which is The MVAR model-based CCL estimation algorithm de- scribed above can be easily extended to estimate the CCLs re- quired to estimate the causal conditional DI, I (X → Y(cid:107)W). required to estimate h (Y(cid:107)W, X). Assuming MVAR model, let Jyy, Kyw, Kyx respectively denote the number of past sam- ples of Y,W, X that influence yn. Then for n = 1, 2,··· , N, the current sample of Y can be expressed as z Jyy(cid:80) Kyw(cid:80) Kyx(cid:80) yn= αjyn−j+ γkwn−k+1+ βlxn−l+1+zn. (14) j=1 k=1 l=1 The only difference with (10) are the extra terms of the time-series W. As a result, the CCL will still be Gaussian distributed with same variance as the distribution in (11) and whose mean contains the extra terms corresponding to the samples of W. The unknown parameters under this model are αj,γk,βl for j=1,··· , Jyy, k = 1,··· , Kyw,l = 1,··· , Kyx and the model orders Jyy, Kyw, Kyx. Maximum likelihood with minimum description length penalty can be used to estimate these parameters similarly. The resulting parameter estimates can then be used to calculate the CCL, which is substituted in (9) to estimate h (Y(cid:107)W, X). Let Jyy and Kyx denote the number of past samples of Y and X that influence the current sample of Y. Then the CCL and , Xn n−Kyx+1 (cid:17) (cid:16) B. Data-driven CCL Estimation P(cid:0)ynYn−1, Xn(cid:1) is same as P (cid:17) (cid:16) can be written as ynYn−1 n−Jyy n−Kyx+1 (cid:16) ,Xn P ynYn−1 n−Jyy (cid:16) (cid:16) Yn Yn−1 n−Jyy n−Jyy P ,Xn ,Xn P = (cid:17) (cid:17) . n−Kyx+1 n−Kyx+1 (cid:17) The joint distribution P of Jyy + 1 and Kyx consecutive samples of Y and X respectively is n−Kyx+1 n−Jyy , Xn Yn 5 Yn , Xn (cid:16) n− Jyy n− Kyx+1 (cid:1) are learned using kernel density estimator [42] with Gaussian kernels. This estimator is implemented in the 'ks' package in R [46]. The true values of (Jyy, Kyx) are not known and should be estimated. The joint density is learned for different values of Jyy and Kyx and the optimal values(cid:0) Jyy, Kyx (cid:17) those that maximize the likelihood. The desired CCL is then estimated by substituting P in (15). The denominator in (15) marginalizes the joint distribution in numerator of (15) over yn. This marginalization is imple- mented by approximating the integral with a Riemann sum of the distribution over a partition of the range of yn. Note that the convergence of the estimated CCL to the true CCL depends on the underlying true data distribution [47]. h (Y(cid:107)X) is obtained by substituting the estimated CCL in (7). can be extended to estimate P(cid:0)ynYn−1, Wn, Xn(cid:1) as well. P(cid:0)ynYn−1,Wn,Xn(cid:1)= Let Jyy, Kyw, Kyx respectively denote the number of past samples of Y, W, X that influence yn. Then The data-driven CCL estimation algorithm described above (cid:17) (cid:17) . (16) (cid:16) (cid:16) ,Wn P n−Kyw +1,Xn n−Kyw +1,Xn n−Kyx+1 n−Kyx+1 Yn Yn−1 n−Jyy n−Jyy ,Wn P The joint distribution in the numerator can be similarly es- timated using kernel density estimator [42] with Gaussian kernels using 'ks' package [46]. Note that the optimal values of the model-orders Jyy, Kyw, Kyx are those that maximize the likelihood. The denominator in (16) is then obtained by marginalizing the distribution in the numerator similarly. The resultant numerator and denominator probabilities are is further substituted in (9) to estimate h (Y(cid:107)X, W). substituted in (16) to estimate P(cid:0)ynYn−1, Wn, Xn(cid:1), which above can be easily modified to estimate P(cid:0)ynYn−1(cid:1), which is required to estimate h (Y). P(cid:0)ynYn−1(cid:1) is obtained from The model-based and data-driven CCL algorithms described either model-based or data-driven CCL by modeling the dependence of the current sample of Y just on its own past samples. I (X → Y) and I (X → Y(cid:107)W) can now be estimated using the estimator proposed in section III. The DI estimator obtained by using the proposed estimator in Theorem. 3.2 with model-based CCL and data-driven CCL estimation algorithms will henceforth be referred to as model- based and data-driven DI estimator respectively. If data is assumed to be drawn from MVAR model with Gaussian white noise, then model-based DI will be referred to as MVAR model-based DI estimator. Note that model-based approach is not restricted to just MVAR models, it is feasible for all those models from which we can estimate the appropriate causal conditional likelihoods parametrically. We focused on MVAR with Gaussian white noise in this paper because ECoG is commonly modeled using this model in connectivity studies [9], [13]. The performance of both the proposed DI estimators on simulated time-series data is demonstrated in the following section. (15) V. PERFORMANCE ON SIMULATED DATA In this section, the performance of the proposed DI esti- mators is demonstrated using simulated data generated from linear (section V-A) five models - two node bidirectional 6 Fig. 1. The true causal connectivity graphs of the simulated data models used to validate the performance of the proposed model-based and data-driven DI estimators. and nonlinear (section V-B) causal network whose true con- nectivity is depicted in Fig. 1a, a two node unidirectional noisy chaotic polynomial (section V-C) causal network whose true connectivity is shown in Fig. 1b, four node linear (sec- tion V-D) and nonlinear (section V-E) causal network whose true connectivity is depicted in Fig. 1c. A directed arrow in Fig. 1 represents a causal connection. The causal connection between two nodes, say from node A to B in Fig. 1c, implies I (A → B) > 0 or equivalently, that the past samples of A have some information about the current sample of B. We also compared the performance of the proposed DI estimators with the standard Granger causality (GC) [12]. GC estimate is obtained from MVGC toolbox [31]. Let us now describe the performance of the proposed DI estimators on the five models considered in detail. A. Two Node Bidirectional Linear Causal Network Consider two time-series X and Y causally connected as shown in Fig. 1a. The time-series Y is generated from yn = β1xn + β2xn−1 + zn, for n = 1, 2,··· , N, (17) where xn and zn are sampled from an i.i.d Gaussian distri- z respectively. The bution with zero mean and variance σ2 samples of X and Z are independent. The true value of the DI between X and Y in both directions is used to benchmark the performance of the proposed model-based and data-driven DI estimators. x, σ2 2 log(cid:0)1 + σ2 Let us first look at the true value of DI for the model by (17) in two special cases. When β1 = 1, β2 = 0, (17) reduces to yn = xn + zn, and it is obvious that both X and Y have equal causal information about each other. It is easy to see that I (X → Y) = I (Y → X) = I (X; Y) = C, where I (X; Y) is the mutual information between X and Y and C = 1 when β1 = 0, β2 = 1 and in this case (17) reduces to yn = xn−1 + zn. In this case, X has causal information about Y, while Y has no causal information about X. More precisely, I (X → Y) = I (X; Y) = C and I (Y → X) = 0. For the remaining case of non-zero β1, β2, the analytical expressions for DI are (cid:1). The other special case occurs x σ2 z I(cid:0)X → Y(cid:1)=1 (cid:16)β1β2σ2 I(cid:0)Y → X(cid:1) = 1 2log σ2 z x (cid:17) (cid:16) 2 log −1 +1 2cosh 1 σ2 1 + β2 σ2 z x (cid:18)(β2 (cid:17) . 2)σ2 1 +β2 2β1β2σ2 x+σ2 z x (cid:19) , (18) The derivation of (18) uses the tridiagonal matrix determinant from [48] and is given in Appendix B. Note from (18) that DI from Y to X does not depend on β2. It is because the (a) β2 = β1 (b) β2 = 1 − β1 Fig. 2. DI estimates and their standard deviation for the two node network (in Fig. 1a) generated from a linear model (17) using analytical expression (18), proposed model-based and data-driven DI estimators for different values of causal strength quantified by (β1, β2). The DI estimates are plotted against β1 with β2 = β1 in Fig. 2a and with β2 = 1 − β1 in Fig. 2b. uncertainty in the current sample of X does not depend on β2, when causally conditioned on the past of X and Y. x = 1, σ2 The DI from X to Y and vice versa is estimated from N = 105 samples of X and Y generated with σ2 z = 1 using the proposed model-based and data-driven DI estimators. The model-based DI estimator assumes that the time-series are modeled by a MVAR model with Gaussian white noise, whereas the data-driven CCL estimator does not impose any model assumptions on the data. Assuming X, Y are from a MVAR process and when xn is included in the past samples of X, Granger causality estimate from X to Y is equal to twice the MVAR model-based DI estimate from X to Y and vice versa [29]. We therefore do not show the GC estimates for linear MVAR models with Gaussian white noise. GC estimates are plotted only for nonlinear simulated models in this paper. Fig. 2 plots directed information values obtained from the analytical expression in (18), I (X → Y) and I (Y → X) from the proposed model-based and data-driven DI estimators for different values of β1 ∈ (0, 1). The corresponding curves are respectively referred to as theoretical, model-based and data-driven. For the model-based and data-driven curves in Fig. 2, multiple datasets of X, Y are generated using different seeds for the random number generator. The mean and the standard deviation of the resultant estimates are plotted in Fig. 2. The average standard deviation across all (β1, β2) in Fig. 2 is about 0.003 and 0.01 for the model-based and data- driven DI estimators respectively. β2 = β1 in Fig. 2a and β2 = 1 − β1 in Fig. 2b. When β1 = β2, a larger β1 implies a stronger causal connection between X and Y and this should result in a larger DI. This expected trend is observed in Fig. 2a. This implies that DI tracks the strength of the causal connection. Also in the corner case of β1 = β2 = 0, DI is zero in both directions as expected. In Fig. 2b, DI estimates in the corner cases of β1 = 0, β2 = 1 and β1 = 1, β2 = 0 match with the analytical expression as expected. Also as β1 increases from 0 to 1, the causal information Y has about X increases, and DI tracks this. This is demonstrated by observing that I (Y → X) increases with β1 in Fig. 2b. Finally, it is clear from Fig. 2 that the model-based estimate matches the correct value of DI estimate from (18) and the data-driven estimator follows the true value of DI. This validates the accuracy of the proposed DI estimators. For this MVAR model with Gaussian XY(a) Two Node Bidirectional NetworkXY(b) Two Node Unidirectional Network(c) Four Node Causal NetworkADCB00.51Connection Strength00.250.5TheoreticalModel-basedData-driven00.51Connection Strength00.20.4TheoreticalModel-basedData-driven white noise, the model-based DI estimator clearly performs better than the data-driven DI estimator and also has a lower run-time. We therefore use the MVAR model-based estimator to estimate DI between data modeled by MVAR processes with Gaussian white noise, instead of using the data-driven estimator. The adaptation of stationary bootstrap algorithm described earlier is used to assess the significance of the inferred causal connections for different values of (β1, β2). We observed that the null hypothesis of no causality from Y to X cannot be rejected for β1 ∈ {0, 0.1} (P-value > δ = 0.05) and can be rejected at all other points (P-value < δ) in Fig. 2. This is not surprising since I (Y → X) is small for β1 ∈ {0, 0.1} and hence did not result in a significant causal connection from Y to X. Similarly, we observed that statistically significant causal connection from X to Y does not exist for β1 = 0, β2 = 0 (P-value > δ) and exits at all other points (P-value < δ) in Fig. 2. This once again confirms our intuition that only large positive values of DI imply a statistically significant causal connection. B. Two Node Bidirectional Nonlinear Causal Network Now, consider time-series X and Y causally connected as shown in Fig. 1a and are generated according to yn = β1x2 n + β2x2 n−1 + zn, for n = 1, 2,··· , N, (19) 1 1 where xn and zn are sampled from an i.i.d Gaussian distribu- z respectively. Also, the tion with zero mean and variance σ2 samples of X and Z are independent. It is very non-trivial to estimate I (X → Y) and I (Y → X) using model-based DI x, σ2 estimator. This is because estimating p(cid:0)xnXn−1 p(cid:0)ynYn−1 (cid:1) and (cid:1) requires essentially inverting the non-linear, non- , Yn 1 Gaussian generative model in (19) and this is very hard even for this simple nonlinear model. These two probability densi- ties are required to estimate h (X(cid:107)Y) and h (Y) respectively. Therefore we only use the proposed data-driven DI estimator to estimate the DI from X to Y and vice versa. However, we can always assume that the data from the model in (19) comes from a MVAR model with Gaussian noise, which is incorrect and estimate DI using the proposed MVAR model- based DI estimator. The resulting DI estimate will be half of the Granger causality estimate between these two time-series, GC (X → Y) and GC (Y → X). Note that GC also assumes the data is generated from a MVAR process even though it is incorrect. We will now compare the performance of data- driven DI and GC estimates on this model. x = 1, σ2 Directed information and Granger causality between X and Y in both directions is estimated from N = 105 samples generated with σ2 z = 1 for different values of (β1, β2) and plotted in Fig. 3. The DI and GC estimates are plotted for β2 = β1 and β2 = 1 − β1 in Fig. 3a and Fig. 3b respectively. For each (β1, β2), multiple datasets of X, Y are generated with different random number generator seeds. The mean and the standard deviation of the resultant data-driven DI and GC estimates are plotted in Fig. 3. The average standard deviation across all (β1, β2) of the data-driven DI and GC estimates is 0.01 and 1.8× 10−5 respectively. In addition, the search space 7 (a) β2 = β1 (b) β2 = 1 − β1 Fig. 3. Data-driven DI and GC estimates, along with standard deviation of the estimates, for the two node network (depicted in Fig. 1a) generated from the nonlinear model (19) for different values of causal strength quantified by (β1, β2). The estimates are plotted against β1 with β2 = β1 in Fig. 3a and with β2 = 1 − β1 in Fig. 3b. of the model order used by the Granger causality estimator is up to 20, i.e, Jyy, Kyx ∈ [1, 20]. In Fig. 3a, I (X → Y) increases with β1 as expected. DI estimates also behave as expected in the corner cases of (β1, β2) = (0, 1) and (1, 0) in Fig. 3b. I (Y → X) increases with β1 as expected. This once again demonstrates that DI tracks the strength of causal connections. On the other hand, Granger causality estimates in both directions are almost zero (of the order of 10−5), indicating that Granger causality cannot detect the causal connections in nonlinear models. The statistical significance of the inferred causal connec- tions by DI and GC estimates for different values of (β1, β2) in Fig. 3 is assessed using the stationary bootstrap algorithm described in section III. Using DI, the null hypothesis of no causality from Y to X cannot be rejected for (β1, β2) ∈ {(0, 0) , (0, 1) (0.1, 0.1) , (0.1, 0.9)} and from X to Y cannot be rejected for (β1, β2) = (0, 0) (P-value > δ = 0.05) in Fig. 3. At all other points in Fig. 3, the null hypothesis of no causality can be rejected (P-value < δ) using DI estimates. This once again confirms our intuition that large values of DI imply a statistically significant causal connection. For GC, the null hypothesis of no causality cannot be rejected at all points in Fig. 3 implying that GC could not find statistically signif- icant causal connections in nonlinear models. This example proves that DI is a more general causal connectivity metric that is not restricted to some particular models. C. Two Node Unidirectional Noisy Chaotic Polynomial Map We now consider two unidirectionally coupled time-series X and Y whose underlying causal connectivity is shown in Fig. 1b. The time-series X and Y are generated from a noisy chaotic polynomial map [49] according to xn = 1.4 − x2 n−1 + 0.3xn−2, (20) yn = 1.4− (βxn−1 + (1 − β) yn−1) yn−1 + 0.3yn−2, where β controls the amount of causal information flowing from X to Y. The initial two samples, x1, x2, y1, y2 are randomly chosen. The two time-series become completely synchronized for β > 0.7. Gaussian i.i.d measurement noise of variance 0.01 is added to both time-series X and Y. For β ∈ [0, 0, 7), strength of the causal connection from X to Y should increase with β and there is no causal connection from Y to X. For β ∈ (0.7, 1], since both time-series are completely -100.51Connection Strength00.250.5^I(X!Y)^I(Y!X)^GC(X!Y)^GC(Y!X)-100.51Connection Strength00.20.4^I(X!Y)^I(Y!X)^GC(X!Y)^GC(Y!X) 8 Fig. 4. Data-driven DI estimates and GC estimates, along with standard de- viation of the estimates, for two node unidirectional network in Fig. 1b gen- erated from noisy chaotic polynomial map (20) for different values of the coupling parameter β. (a) Model-based Estimator (b) Data-driven Estimator Fig. 5. The causal network along with connection strengths between the four MVAR processes simulated from (21) estimated by the MVAR model-based DI and the data- driven DI estimators. The true causal connectivity graph between these four time-series is depicted in Fig. 1c. It is clear that both DI estimators correctly infer the underlying causal network. (a) Model-based Estimator (b) Data-driven Estimator Fig. 6. The causal network along with connection strengths between the four time-series simulated from (22) estimated by the MVAR model-based DI and the data-driven DI es- timators. The true causal connectivity graph between these four time-series is depicted in Fig. 1c. It is clear that unlike MVAR model-based estimator, the data-driven estimator correctly infers the underlying causal connectivity graph. synchronized and because of the measurement noise, there is a non-zero equally strong causal connection in both directions. In the absence of measurement noise for β ∈ (0.7, 1), xn = yn leading to causal conditional entropy estimate of negative infinity and a DI estimate of infinity. The intuition behind this is that once the past of X is known, there is no uncertainty left in Y. On the other hand, GC estimates in the synchronized range will be close to zero because the past of X used by GC (unlike DI, GC does not include xn in the past of X) does not contain any predictive information about yn resulting in a GC estimate of zero from X to Y. Note that it is very non-trivial to apply model-based DI on this model because of the same reasons outlined in the previous simulated nonlinear model. We therefore only compare the performance of data-driven DI and GC estimates on this model. DI and GC in both directions is estimated from N = 105 samples of X and Y (after discarding the initial transient points) for different values of β ∈ [0, 1] and plotted in Fig. 4. For each β, the time-series are generated from (20) using different seeds of the random number generator. The mean and the standard deviation of the resulting data-driven DI and GC estimates are plotted in Fig. 4. The average standard deviation across all β for the data-driven DI and GC estimates is 0.03 and 0.001 respectively. The standard deviation was largest at β = 0.7, implying that it is very hard to estimate at the bound- ary before and after complete synchronization. In addition, the search space of the model order used by the Granger causality estimator is up to 20, i.e, Jyy, Kyx ∈ [1, 20]. The DI estimate is obtained by subtracting two non-negative numbers and it can sometimes be a small negative number because of the inaccuracies in estimation algorithms or insufficient data or violation of stationarity assumptions [23] and in those cases, we reset the DI estimate to be zero. For instance, the largest negative DI estimate we obtained for this model is −0.06 from Y to X at β = 0.6 and we reset this estimate to 0. It is clear from Fig. 4 that DI estimates behave as expected. DI from X to Y increase as β goes from 0 to 1. On the other hand, the DI estimates from Y to X are very small numbers for β < 0.7 and then there is a sudden jump in this estimate after β > 0.7. This jump is because the time-series get synchronized for β > 0.7. On the other hand, GC estimates in both directions are small positive numbers (when compared to DI estimates) for the whole range and become equal in value in the synchronized range of β > 0.7. The statistical significance of the causal connections in- ferred by DI and GC estimates is assessed using the adaption of stationary bootstrap. The null hypothesis of no causality using DI estimates from Y to X cannot be rejected for β < 0.7 and cannot be rejected for the connection from X to Y for β < 0.1. This implies DI correctly identifies the presence of causal connection from X to Y for all β ≥ 0.1 and the absence of causal connection from Y to X for β < 0.7. It can also differentiate causally independent time-series (β = 0) and completely identical time series (β ∈ (0.7, 1]). On the other hand, the null hypothesis of no causality cannot be rejected only for β = 0 using GC estimates. This implies GC identifies the presence of a causal connection in both directions for all non-zero β, which is incorrect. This example also shows DI correctly infers causal connectivity from nonlinear models. D. Four Node Linear Causal Network Now, consider the four node causal network depicted in Fig. 1c. The four time series A, B, C and D are generated according to bn = an−1 + an−2 + zb n, dn = an−2 + zd n, n, zc n and zd cn = bn−1 + zc n, for n = 1, 2,··· , N, (21) where an, zb n are sampled from an i.i.d Gaus- sian distribution with zero mean and unit variance. In this network, A influences C indirectly through B. This is an example of an 'indirect' causal connection, in contrast with the connection from A to B, which is a 'direct' causal connection. DI estimate between pairs of time-series cannot differentiate between 'direct' and 'indirect' causal connections [17]. For instance, the DI estimate from A to C is positive, even though A does not directly influence C, but causally influences C via B. A thorough discussion on the direct and indirect influences for point processes is in [17] and is directly applicable here. Following the approach taken in [29], [50], influence from A to C is non-zero, if and only if I (A → C(cid:107)B, D, ) > 0. However, estimating the causally conditioned DI when the number of channels recorded from is large (of the order of hundred's) is difficult because of the curse of dimensionality [43]. To overcome this, the pairwise DI is first estimated between all the 'direct' causal 00.50.71Connection Strength0 0.30.6ADCBADCBADCBADCB pairs of channels. The indirect influences are then resolved by first estimating only the required causal DI between two processes, conditioned on one more process. Then if required, the causal DI between two processes, conditioned on two more processes, is estimated and so on. The termination condition is determined by the desired degree of 'directness' in the inferred causal network. In this simulated example, we are interested in recovering the true 'direct' causal network depicted in Fig. 1c. To infer the true causal network, DI is estimated between these four time-series using both MVAR model-based and data-driven DI estimators. Model-based DI estimator assumes the data is generated from a linear causal MVAR model, whereas data-driven DI estimator does not impose any para- metric model assumptions on the data. The data is generated from (21) using 20 different seeds to generate the Gaussian noise and the resultant estimates are averaged. We will first describe the performance using model-based DI estimator. Model-based DI estimator is used to estimate the pair- wise DI between all pairs of these four nodes, resulting a 4 × 4 matrix with zeros on the diagonal. We found that I (A → B) = 0.485 ± 0.009, I (A → C) = 0.314 ± 0.009 and I (B → C) = 0.658 ± 0.009. To determine if there is an indirect causal connection from A to C or from B to C, we estimated I (A → C(cid:107)B) and I (B → C(cid:107)A) using the model-based causally conditioned DI estimator described in section III, IV-A. We found that I (A → C(cid:107)B) = 0 and I (B → C(cid:107)A) = 0.344 ± 0.009. Therefore, A to C is an 'indirect' connection via B. Causally conditional DIs are estimated till the network is completely resolved and free of any indirect influences. The estimated causal network along with the strength and the standard deviation of the estimated causal connections is depicted in Fig. 5a. It is clear from Fig. 5a and Fig. 1c that model-based DI estimator infers the true causal network correctly. We now use the data-driven DI estimator to infer the true causal network. The pairwise DI is estimated between all pairs of these four nodes using the data-driven estimator, resulting in a 4 × 4 matrix with zeros on the diagonal. Using this DI estimator, we find that I (A → B) = 0.468 ± 0.009, I (A → C) = 0.296± 0.004and I (B → C) = 0.648± 0.008. To identify the presence of any indirect connections, we estimated I (A → C(cid:107)B) and I (B → C(cid:107)A) using the model-based causally conditioned DI estimator described in section III, IV-B. We found that I (A → C(cid:107)B) = 0 and I (B → C(cid:107)A) = 0.273 ± 0.009. Therefore, A to C is an 'indirect' connection via B. This procedure is continued to identify and remove all indirect causal connections. The resultant estimated direct causal network is depicted in Fig. 5b. It is clear that data-driven DI also recovers the true network correctly. Moreover, it is clear from Fig. 5 that for this model, both model-based and data-driven DI estimators correctly infer the underlying causal network, which is not surprising since the underlying model is a linear MVAR model. E. Four Node Nonlinear Causal Network We now use a nonlinear model to generate the four time- series A, B, C and D whose underlying causal connectivity 9 graph is depicted in Fig. 1c. N samples from the four time- series are generated according to bn = a2 n−2 + zb n−1 + a2 n, dn = an−2 + zd n, cn = bn−1 + zc n, for n = 1, 2,··· , N, (22) n, zc n and zd where an, zb n are sampled from an i.i.d Gaussian distribution with zero mean and unit variance. The only difference with the model in section V-D is that the causal connection from A to B is now nonlinear. First, we infer the true causal connectivity for this model using the MVAR model-based DI estimator. This DI estimator assumes that the data is drawn from a linear MVAR model, which is not true for this model. It is clear from (22) that the time-series B is not generated from a linear MVAR model. Pairwise DI is estimated using this model between all pairs of these four time-series resulting in a 4 × 4 matrix with zeros on the diagonal. The only significant causal connections estimated by the model-based DI estimator are from B to C and from A to D. This process is repeated for data generated using 20 different seeds and the resultant DI estimates are averaged. We find that I (B → C) = 0.847 ± 0.013 and I (A → D) = 0.348 ± 0.008. It is also clear that there are no indirect connections to resolve in this case. The underlying causal connectivity graph estimated by the model-based DI estimator is depicted in Fig. 6a. It is clear from this figure that model-based DI estimator could not recover this true network correctly. This is not surprising since the MVAR model- based estimator can only identify linear causal connections and cannot identify the nonlinear causal connections. As result, the connection from A to B is not identified by the model-based DI estimator. We now use data-driven DI estimator to infer the causal connectivity from the simulated data. The pairwise DI is estimated between all pairs of these four nodes using the data-driven estimator, resulting in a 4 × 4 matrix with zeros on the diagonal. In contrast to the model-based DI estima- tor, we find that DI from A to B estimated using data- driven DI is nonzero. Specifically, we find that I (A → B) = 0.433 ± 0.011. In addition, we also find that I (A → C) = 0.320 ± 0.010 and I (B → C) = 0.753 ± 0.009. To eliminate indirect causal connections, we estimated I (A → C(cid:107)B) = 0 and I (B → C(cid:107)A) = 0.262 ± 0.037. Therefore, A to C is an 'indirect' connection via B. This procedure is continued to identify and remove all indirect causal connections. The resultant estimated direct causal network is depicted in Fig. 6b. It is clear that data-driven DI estimator recovers the true network correctly, while the model-based DI estimator could not infer the true causal network correctly. The five diverse simulated models considered in this section demonstrate that the DI correctly infers the presence and tracks the strength of a causal connection - large values of DI imply a strong causal connection and vice versa. Using stationary bootstrap, we also showed that only large positive DI estimates correspond to statistically significant causal connections. We also observed that model-based DI estimator cannot identify nonlinear causal connections, whereas data-driven DI estima- tor can correctly identify both linear and nonlinear causal 10 CLINICAL DETAILS OF THE PATIENTS ANALYZED. TABLE I P1 Patient ID Age/Sex Syndrome 20/M Nonlesional temporal Lesional 60/M temporal 29/M Nonlesional temporal 37/M Nonlesional Lesional temporal 20/F P4 P2 P3 P5 extratemporal SPS+CPS Seizure Type CPS CPS CPS Electrode Type Surgery Outcome of Surgery D Right TL Class I Selective Left HC Class II G+D Right TL Class II D G Right OC Class III CPS G+D Left TL Class I CPS - complex partial seizures, SPS - simple partial seizures. D - depth electrodes, G - subdural grid electrodes. TL - temporal lobectomy, HC - hippocampectomy, OC - occipital corticosectomy. The outcomes are in Engel epilepsy surgery outcome scale. "Class I - free of disabling seizures, class II - Almost seizure-free, class III - worthwhile improvement, class IV - no worthwhile improvement" [52]. connections. We now use both the MVAR model-based and data-driven DI estimators to infer the causal connectivity graph from ECoG data in epileptic patients. We only consider the large DI estimates (large compared to the rest of the causal connectivity graph) since they only imply a significant causal connection. We propose a model-based and a data-driven SOZ identification algorithm in the following section. VI. SEIZURE ONSET ZONE IDENTIFICATION ALGORITHMS Seizure onset zone (SOZ) is defined as the regions of the brain that initiate seizures [8]. The current clinical standard is for neurologists to identify SOZ from visual analysis of the ECoG data. The SOZ identified in this way is removed during resective surgery. However, visual analysis is time consuming, subjective and potentially unreliable [36], [51]. We propose two computationally derived SOZ identification algorithms - model-based and data-driven SOZ identification algorithms. We identified the SOZ in five patients with epilepsy using these two algorithms and compared their performance with visual analysis by the neurologist. A. Clinical ECoG Data The five patients analyzed here were all managed and treated by our physician coauthors. The clinical details of these patients are summarized in Table I. Three seizure records each from patients P1, P2 and P5, two from patient P3 and one from patient P4 were analyzed. Each seizure record was approximately 10 minutes long and contained one seizure. Each seizure on average lasted for a minute and was roughly in the middle of the seizure record. The seizure start time was identified by the neurologist. Each electrode records the voltage waveform at a sampling frequency of 1 KHz. The number of electrodes in these five patients varied from 120 to 150. Electrodes with artifacts likely due to either loose contacts, patient movement or excessive line noise were not included in the analysis. Fig. 7. A 30s snapshot of ECoG signals from the 30 high energy channels of P1. The seizure start time, represented by a vertical solid black line, is identified by neurologist. Causal connectivity is estimated from this entire 30s window for this seizure record. B. Proposed SOZ Identification Algorithms The first stage of the proposed SOZ identification algorithm is an energy detector which selects only M channels out of all ECoG channels for further analysis. The main objective of this stage is to reduce the computational complexity of the proposed algorithms. The energy is l2-norm of the ECoG signal computed from a window around the start of seizures containing preictal and ictal recordings. Any channel involved in seizure onset is expected to have interictal spikes before the seizure starts and/or have high amplitude low-frequency ictal activity once the seizure is fully developed, both of which will increase the energy in the selected time-window. The time- window was selected to be long enough to capture both spiking and large ECoG amplitudes during seizures. The second stage consisted of estimating the causal connectivity between every pair of M channels selected in the first stage to form a M × M causal connectivity matrix. The causal connectivity was estimated from a shorter time-window around the seizure start time, since we are interested in estimating the seizure onset electrodes. The following subsections describe the remaining stages of the two proposed SOZ identification algorithms. 1) Model-based SOZ Identification Algorithm: In this ap- proach, ECoG data is assumed to be derived from a MVAR process with Gaussian white noise. This is a very common assumption imposed to estimate causal connectivity between ECoG data [9], [13]. The MVAR model-based DI estimator is used to infer the causal connectivity between the selected M high energy channels. The causal connectivity estimated using this approach only represents the linear causal interactions between the ECoG channels. However it is widely believed that seizures are highly non-linear phenomenon during which SOZ drives the rest of the network into a hypersynchronous state [4], [8], [16]. As a result, we expect the seizure onset electrodes in the causal connectivity graph to be isolated, since model-based approach can only capture linear causal interactions. The proposed model-based algorithm therefore identifies the nodes in the causal connectivity graph with zero degree (threshold was set to select only the strongest 10% connections) as the estimated SOZ. If a patient had multiple seizures, the electrodes identified across all seizures in that patient form the estimated SOZ for that patient. 2412512612713020101Time (s)Channel Index 11 (a) From Patient P1 (b) From Patient P3 Fig. 9. Normalized net outward flow from the ECoG electrodes with positive net information outflow using data-driven SOZ identification algorithm. activity (150 past samples at Fs = 1KHz) at this channel and other channels. This corresponds to restricting the model order Jyy, Kyx search space to [1, 150] for the MVAR model-based DI estimator. In addition, we need to capture the connectivity just before and just after a seizure starts to estimate the SOZ. Therefore, we used ECoG data from a 30s window (3 × 104 data points) that begins 20s before the start of the seizure to be stationary. The same window was used for the data-driven estimator as well. In addition, the past activity was down- sampled by a factor of 50 for the data-driven estimator to restrict the Jyy, Kyx search space to [1, 4] and also reduce its computational complexity (i.e. the past activity of channel X can include {xn, xn−50, xn−100, xn−150}). The exact values of these parameters is not crucial as the algorithms seem to be fairly robust to changes in these parameters. Consider the second seizure record of patient P1. The energy detector selected 30 high energy channels. Fig. 7 shows the recordings from these channels in the 30s window in which causal connectivity graph is inferred. The inferred graph by model-based and data-driven approaches is shown in Fig. 8. The weighted adjacency matrix of the inferred causal connectivity graphs, whose (i, j)th element is the DI estimate from channel i to j for i, j ∈ [1, 30], is plotted in Fig. 8 using a image plot. It is clear from this figure that the mean strength of the DI estimates using model-based approach is smaller than using data-driven approach (colorbar ranges are different in the two sub-figures). We observed this across all the twelve seizures analyzed. This indicates that data- driven DI captured more causal information on average than model-based DI, implying that non-linear causal interactions are stronger around the beginning of a seizure. The nodes with zero degree in the causal connectivity graphs from each seizure in a patient are identified as the SOZ by the model- based algorithm. The zero degree criterion used by model- based algorithm is counterintuitive, since we except the SOZ to drive the network to seizure state and not be weakly connected. On the other hand, the data-driven algorithm selects electrodes with large net outflows, which is very intuitive. The data- driven algorithm computed the normalized net outward flow for each node using (24). Fig. 9a plots the Φ for all electrodes with positive net outward flows in patient P1. The electrodes with Φ > 5% are the estimated SOZ for this patient P1 using data-driven algorithm. Table II summarizes the results from our analysis. The first column in Table II identifies the patient ID and the number (a) From model-based algorithm (b) From data-driven algorithm Fig. 8. Causal connectivity between 30 high energy channels estimated from ECoG data between 241s and 271s from the second seizure of P1. The channel indices with bluish rows and bluish columns (correspond to low DI estimates) in Fig. 8a correspond to isolated nodes and are the estimated SOZ using model-based algorithm. The corresponding channels in Fig. 8b have large net- outflows of information and are the estimated SOZ from data-driven algorithm. 2) Data-driven SOZ Identification Algorithm: In this al- gorithm, no parametric model assumptions were imposed on ECoG data. The causal connectivity between the M high energy channels selected in the first stage was inferred using the data-driven DI estimator. This estimator inferred both linear and nonlinear causal interactions between channels. Intuitively, activity at the SOZ electrodes drives the activity at the other electrodes into a hypersynchronous state via linear and nonlinear causal interactions [16]. We therefore expect the SOZ electrodes to act as sources (with strong outgoing and weak incoming causal connections) in the causal connectivity graph inferred around the seizure start time using data-driven DI. As a result, the SOZ nodes in the causal connectivity graph are expected to have large net-outward flow of information. The data-driven SOZ identification algorithm quantifies this intuition to estimate SOZ. The net-outward flow (Φ) of causal information from an electrode i is calculated using {I(i → j) − I(j → i)} . M(cid:80) Φ(i) = (23) j=1,j(cid:54)=i If a patient had multiple seizures, the net outward flow of an electrode is the average net outward flow of that electrode across all seizures recorded in that patient. Then the normal- ized net outward flow ( Φ) of the electrode i is given by Φ(i) = 100 × Φ(i)(cid:80) Φ(j) . (24) j:Φ(j)>0 The electrodes with Φ > 5% are considered to have significant net outward flow of information in the causal connectivity graph and are identified as the seizure onset electrodes for that patient by the data-driven SOZ identification algorithm. C. Performance of Proposed SOZ Identification Algorithms The energy detector selected the top M = 30 channels with the largest energy computed from a 100s window com- prising of 50s of activity immediately before and after the seizure starts. The causal connectivity graph between these high energy channels is then estimated using model-based and data-driven DI estimators from a 30s window that begins 20s before the seizure start time and ends 10s into the start of the seizure. We assumed that the current activity at an ECoG channel does not depend on more than 150ms of past Channel Index1 1530Channel Index1 153000.10.20.30.40.5Channel Index1 1530Channel Index1 15300 0.10.94.6 12 of seizures analyzed for that patient. The second, third and fourth columns in Table II list the SOZ identified across all the five patients using model-based, data-driven algorithms and visual analysis respectively. We observed that all the channels identified as SOZ by visual analysis, except AST 2 in one seizure of P5, are included in the 30 high energy channels selected from each seizure by the energy detector in the first stage. The top 30 channels selected from two seizures in patient P5 contained AST 2, but the 30 channels picked from the third seizure did not contain AST 2. The normalized net outflow Φ from AST 2 electrode for patient P5 using data-driven algorithm was 1% and hence this electrode was not identified as SOZ (note that Φ has to exceed 5% to be selected as SOZ). Expect for this one region, it is clear from this table that the data-driven algorithm identifies all the regions identified by the neurologist, whereas the model- based algorithm misses some regions (for instance, RAMY electrodes in P1, TP and AST electrodes in P5). Also, the model-based algorithm incorrectly identified lateral temporal (LT) electrodes as SOZ in patient P3, whereas data-driven algorithm correctly identified posterior depth (PD) electrodes in hippocampus as SOZ. Except in P3 and P4, both algorithms do not have any false positives. The false positives in P4 could be because only one seizure was analyzed in this patient. Another advantage of the data-driven SOZ identification algorithm over model-based algorithm and analysis by the neurologist is that Φ could be used as a quantitative metric to rank the electrodes in the decreasing order of clinical relevance. Fig. 9 plots the Φ of all electrodes with positive net outward flows in patients P1 and P3. Using our metric Φ, it is clear from Fig. 9b that electrodes PD4, PD5 contribute much more in generating and spreading seizures than LF28 and LP4 electrodes even though Φ exceeds the chosen threshold at all these four electrodes. Depending on the significance level (5% is used here), the set of selected SOZ electrodes varies. We observed in all five patients that the electrodes with the highest Φ values were always the same as the ones identified by the neurologist. Visual analysis by the neurol- ogist can only give qualitative information about the SOZ and cannot give quantitative information like the proposed data-driven SOZ identification algorithm. In addition, data- driven algorithm can also differentiate between electrodes in close proximity - for example, Φ is negative for RPH2 electrode in P1 even though Φ is positive for both RPH3 and RPH4 (refer to Fig. 9a). The increased spatial-specificity provided by our data-driven algorithm could be relevant for next generation epilepsy treatments [5]. The main advantage of the model-based algorithm over data-driven one is its lower computational complexity. However, this is less critical with today's powerful computers. To summarize, data-driven SOZ identification algorithm outperforms model-based algorithm and provides more interpretable results. VII. DISCUSSION AND CONCLUSIONS An almost surely convergent MVAR model-based and data- driven estimators for DI are introduced in this paper. Linear causal interactions between two time-series can be quantified SEIZURE ONSET ZONE IDENTIFIED FROM THE PROPOSED ALGORITHMS AND THE VISUAL ANALYSIS BY NEUROLOGIST. TABLE II Patient - # of Seizures P1 - 3 P2 - 3 P3 - 2 P4 - 1 P5 - 3 Model-based Algorithm Data-driven Algorithm RAH 1-3, RPH 2-4 RAH 1-2, RPH 4, LAH 2-4, LPH 2 LAH 2-4, LPH 1-2 RAMY 2-3 LT 1-3, 10 LO 3, 14, 15, 25, LO 12, 13, PST 3, PST 1, MOG 27 MST 1, 2, HD 1 PD 4-5, LF 28, LP 4 LO 3, 14, 15, 12, PST 1, MOG 23, SOG 21, 36 MST 1, TP 1, HD 1 Visual Analysis RAH 1-3, RPH 2-4, RAMY 2-3 LAH 2-4, LPH 1-2 PD 3-5 LO 3, 14, 15, LO 25, PST 3 MST 1, 2, TP 1, HD 1-3, AST 2 The label of an ECoG electrode comprises of an abbreviation of the brain region it is implanted in and a number. For depth electrodes, smallest number is assigned to deepest electrode from scalp. For instance, RAH1 - deepest electrode contact in depth electrode in right anterior hippocampus and LO3 - third electrode contact in subdural grid electrode over lateral occipital lobe. RPH - right posterior hippocampus, RAMY - right amygdala, LF - lateral frontal, LP - lateral parietal, LT - lateral temporal, PD - posterior hippocampal depth, MOG - medial occipital grid, SOG - sub-occipital grid, PST - posterior sub-temporal, MST - mid-subtemporal lobe AST - anterior sub-temporal lobe, TP - temporo-polar, HD - hippocampal depth. using MVAR model-based DI estimator, whereas both linear and nonlinear causal interactions are quantified by data-driven DI estimator. The resultant DI estimates can be used to infer whether the data has (1) linear causal interactions or (2) both linear and nonlinear causal interactions. If the MVAR model- based DI estimate is comparable in value to data-driven DI estimate, then the interaction is predominantly linear. This is not feasible with existing metrics because they can be split into two non-overlapping groups - the first group only detects linear causal interactions (e.g., Granger causality, partial directed coherence), whereas the second group detects both linear and nonlinear causal interactions (e.g., transfer entropy). The DI estimators proposed in this paper can be automatically adapted to other types of electrophysiological data like EEG to learn the causal connectivity. Data-driven DI estimator seems to be more appropriate than model-based DI estimator if the underlying data distribution is not known, which is the case with most real data. The main challenge with data-driven DI estimator is estimating the causal conditional likelihood nonparametrically and its computational complexity. We used kernel density estimators in this paper to estimate causal conditional likelihood. Ker- nel density estimators are asymptotically optimal [43]. Their bias decreases with increasing number of data samples and complexity increases with the dimensionality of the data, just like other nonparametric estimators. Even though we selected optimal bandwidth using smoothed cross-validation to minimize the asymptotic mean integrated squared error, several other criteria could also be used [43], [46]. In addition, data-driven entropy estimators based on adaptive partitioning, nearest neighbors and m-spacing algorithms [24], [53] can also be used to estimate DI nonparametrically. Another approach to estimate DI nonparametrically is to extend the universal DI estimator proposed for discrete-valued signals in [23] to continuous-valued ECoG signals. Future work should also include developing approximate data-driven DI estimators to 13 Fig. 10. Causal connectivity between 30 high energy channels depicted in Fig. 7 estimated using PDC. (a) From a segment before the seizure (c) From a segment after the seizure Fig. 11. Causal connectivity between the 30 high energy channels from the second seizure of patient P1 estimated using data-driven DI estimator from ECoG data in three segments - one before seizure (181s -211s), one during seizure (261s - 291s) and one after seizure (361s - 391s). This seizure starts at 261s and ends at 350s. (b) From a segment during the seizure further reduce computational complexity. Directed information was used in this paper to estimate causal connectivity between ECoG channels. The causal con- nection identified between two channels could be due to the effect of activity at other spatial locations in the brain. If an ECoG electrode was implanted at these other locations, causally conditioned DI can be used to remove their influ- ence. This was demonstrated using the four node examples in section V. On the other hand, if ECoG activity is not recorded from these locations, then removing the effects of these hidden nodes on the inferred causal connectivity is a very hard problem in general. Future work should look into the sensitivity of DI to volume conduction effects when compared with synchronization metrics like phase lag index [54]. DI estimators proposed in this paper do not quantify the amount of causal information between time-series at each frequency, unlike partial directed coherence (PDC) or directed transfer function (DTF). However, the advantage of DI is that data-driven DI estimator can detect nonlinear causal interactions, which PDC or DTF cannot detect. Metrics based on PDC, DTF assume the data is drawn from a MVAR model and can only detect linear causal interactions (similar to MVAR model-based DI estimator proposed in this paper). To demonstrate this, we estimated the causal connectivity graph between the 30 channels depicted in Fig. 7 by PDC using eMVAR toolbox [55]. The resultant 30 × 30 causal connectivity matrix is plotted in Fig. 10, in which (i, j) element corresponds to the maximum value of PDC from channel i to channel j. Note that causal connectivity estimates from the proposed DI estimators for the same data is plotted in Fig. 8. Comparing Fig. 10 with Fig. 8b, it is clear that net outflow from the SOZ electrodes is not large in PDC when compared to data-driven DI. This implies unlike data-driven DI estimator, PDC cannot capture nonlinear causal interactions. We also proposed model-based and data-driven algorithms to identify the SOZ. The first stage of both these algorithms is an energy detector. The chosen electrodes from the first stage turned out to have large overlap (more than half) across multiple seizures within a patient. All electrodes with low rhythmic gamma activity in SOZ were selected by the energy detector in all the patients analyzed. Note that other criteria could also be used instead of energy detector. In particular, we experimented with selecting channels displaying strong high- frequency activity around the seizure start time (since channels involved in seizure onset display strong high-frequency activ- ity around the beginning of a seizure that typically develops into high amplitude low-frequency activity). The time-window used to estimate the high-frequency activity should be of much smaller length than the one used with energy detector, because the seizures typically display low amplitude rhythmic high-frequency oscillations only for a very short duration. The resulting performance with energy detector or the high- frequency activity detector was similar. We therefore presented the results only with the energy detector in this paper. The causal connectivity graphs between the selected high energy channels estimated using MVAR model-based DI and data-driven DI from same time-window are not the same, since both estimators capture different causal interactions in the data - model-based captures linear interactions, whereas data- driven captures both linear and nonlinear causal interactions. Therefore the criterion used to estimate SOZ from the causal connectivity graph was different for the two algorithms. In the model-based approach, the SOZ nodes are isolated since they drive the other brain regions into a seizure through nonlinear interactions (which are not captured by model-based DI estimator). Similar results were reported in other studies using linear metrics [56], [57]. It is reported in [57] that SOZ electrodes form an isolated focus using symmetric coherence metric that captures linear interactions. On the other hand in causal connectivity graphs estimated by data-driven DI, the outgoing and incoming edges from SOZ electrodes have large and small DI estimates respectively (refer to Fig. 9). This is in accordance with our intuition that the SOZ drives the seizure activity [4], [8], [16]. Also metrics closely related to net outward flow were used in [35] to infer SOZ using transfer entropy (which detects nonlinear interactions) by analyzing hours of ECoG recordings (here we are only using recordings from a 30s window). A key advantage of ECoG recordings over other neu- roimaging techniques is its good temporal resolution. The DI estimators proposed in this paper can be applied to ECoG recordings from different windows to learn the spatiotemporal changes in causal connectivity networks during the course of a seizure. The causal connectivity before, during and after the second seizure of patient P1 estimated using data-driven DI estimator from three 30s long windows is shown in Fig. 11. It is clear from Fig. 11 that SOZ electrodes (corresponding to rows with more red color or large DI values in Fig. 11b) have large net outflows during seizure when compared with before and after seizure (same rows have more blue color or Channel Index1 1530Channel Index1 15300 0.10.20.30.4Channel Index1 1530Channel Index1 153000.71.42.12.8Channel Index1 1530Channel Index1 15300.10.41.86.3Channel Index1 1530Channel Index1 153000.61.21.8 14 smaller DI value in Fig. 11a , 11c). We are extending this analysis to infer seizure mechanisms by examining the changes in causal connectivity estimated from preictal, ictal, postictal periods when compared with interictal periods. This is the subject of our current and future work [58]. The results from this analysis potentially could improve our understanding of seizure mechanisms and lead to the development of novel non- surgical treatments for epilepsy. VIII. ACKNOWLEDGMENTS The authors wish to thank Sugnaya Karunakaran for care- fully reading the manuscript. PROOF OF CAUSAL CONDITIONAL ENTROPY ESTIMATOR APPENDIX A A. Proof of Lemma 3.1 1 (cid:1) (25) n−K+1 n−J , Xn (cid:1) ≥ ······ First, we will prove the existence of h (Y(cid:107)X). Since conditioning reduces differential entropy, we have increasing sequence that is upper bounded by h (y1). Also let l = max (J + 1, K). Then for n ≥ l, 1, X2 1 n−K+1 n−J , Xn (cid:1) ≥ h(cid:0)y2Y1 h (y1) ≥ h(cid:0)y1X1 ≥ ··· ≥ h(cid:0)ynYn−1 Therefore the sequence h(cid:0)ynYn−1 h(cid:0)ynYn−1 (cid:1) = h(cid:0)ynYn−1 = h(cid:0)ylYl−1 the sequence h(cid:0)ynYn−1 (cid:1). Let an = h(cid:0)ynYn−1 h(cid:0)ylYl−1 N h(cid:0)YN(cid:107)XN(cid:1) = 1 N(cid:80) (cid:1) is a non- (cid:1) (cid:1) , (cid:1) is lower bounded by (cid:1) and bN = where (26) is from the Markovian assumption and (27) is from the stationarity assumption. Note that (27) also implies N→∞ aN exists, from Cesaro mean theorem [30] we have h (Y(cid:107)X) = lim N→∞ bN also exists. The above proof can be easily modified to prove h (Y) exists. Therefore I (X → Y) = h (Y)− h (Y(cid:107)X) also exists. n−J , Xn l−J , Xl an. Since the lim n−K+1 l−K+1 n−J , Xn (26) (27) l−J , Xl n−K+1 l−K+1 , Xn 1 , Xn 1 n=1 N 1 1 1 1 B. Proof of Lemma 3.2 N h(cid:0)YN(cid:107)XN(cid:1) = 1 N(cid:80) h(cid:0)ynYn−1 N(cid:80) E(cid:2)− log P(cid:0)ylYl−1 n=1 = 1 N = E(cid:2)− log P(cid:0)ylYl−1 n=1 l−J , Xl N n−J , Xn n−K+1 (cid:1)(cid:3) , l−J , Xl l−K+1 l−K+1 (cid:1) (cid:1)(cid:3) (28) (29) where (28) is from chain rule and Markovian assumption, and (29) is due to stationarity. (cid:16) (cid:17) C. Proof of Theorem 3.1 n−K+1 be a fixed function over the states of the Markov chain = −log P(cid:0)ynYn−1 (cid:1) (cid:1). From the strong law of large numbers for (cid:0)Yn Let gJ,K n−J,Xn n−J,Xn n−K+1) Yn n−J,Xn n−K+1 Markov chains [39] which states that for a fixed function g (.) over the states of the Markov chain, the sample mean will N(cid:80) 1 N n=1 gJ,K We also have h(Y(cid:107)X)= lim N→∞ almost surely converge to the expected value as N → ∞, we have, n−K+1 n−J,Xn (cid:0)Yn (cid:1) a.s.−−→E(cid:2)gJ,K (cid:0)Yl Nh(cid:0)YN(cid:107)XN(cid:1)= lim E(cid:2)gJ,K (cid:0)Yl = E(cid:2)gJ,K (cid:0)Yl (cid:1)(cid:3) , N(cid:80) N→∞ l−K+1 1 l−J,Xl l−K+1 (cid:1)(cid:3) . (30) (cid:1)(cid:3) (31) (cid:17) a.s.−−→ h (Y(cid:107)X) . l−K+1 l−J,Xl l−J , Xl (32) where (31) is from Lemma. 3.2. We have from (30), (32), h (Y(cid:107)X) = 1 (cid:16) Yn n−J , Xn n−(K−1) gJ,K N n=1 APPENDIX B DERIVATION OF DI FOR LINEAR TWO NODE NETWORK Consider the MVAR model in section V-A, described by (17). Here we will derive the DI in both directions between time-series X and Y for non-zero β1, β2. Appendix. B-C considers the case when (β1, β2) ∈ {(1, 0) , (1, 0)}. A. DI from X to Y For the system described by (17), the causal conditional entropy h (Y(cid:107)X) is given by N h(cid:0)YN(cid:107)XN(cid:1) = 1 2 log(cid:0)2πeσ2 1 z (cid:1) , h (Y(cid:107)X) = lim N→∞ (33) where γ = (cid:0)β2 because conditioned on (xn, xn−1), the only uncertainty in yn z which is is due to the i.i.d Gaussian noise Z of variance σ2 independent of X. Now, from (17), we have (y1, y2,··· , yN )T ∼ N (0, ΣN ), x. MN is a tridiagonal where ΣN = δMN with δ = β1β2σ2 matrix whose main diagonal elements are D and non-zero diagonal below and above the main diagonal are all 1. D = γ δ , z. Upon further simplification using the tridiagonal matrix determinant from [48], we have ΣN = δN sinh((N +1)λ) The unconditioned entropy of Y is now given by 2 log(2πeδ)+1 =1 2 λ, −1(cid:16)D , where λ = cosh (cid:1) σ2 x + σ2 1 + β2 2 1 2Nlog (cid:17) (34) (35) sinh λ h (Y)= lim N→∞ 2 . obtained by expanding the hyperbolic sinh function in the determinant ΣN in terms of exponentials and some basic algebraic manipulations. Now, from (33) and (35), we have I (X → Y) = 1 (cid:18) (β2 (cid:19) x+σ2 + 1 −1 x z . 2 cosh 2)σ2 1 +β2 2β1β2σ2 x 2 log σ2 z (cid:16) (2πe)NΣN(cid:17) (cid:16)β1β2σ2 (cid:17) B. DI from Y to X The causal conditional entropy, h (X(cid:107)Y) is given by h (X(cid:107)Y) = lim N→∞ h (xnxn−1, yn) N(cid:80) 1 N n=1 (36) (37) N(cid:80) N(cid:80) n=1 n=1 {h (xn, yn, xn−1) − h (xn−1, yn)} (cid:8) 1 2 log (2πeΦ1) − 1 2 log (2πeΦ2)(cid:9) = lim N→∞ 1 N 1 N (cid:16) = lim N→∞ = 1 2 log (cid:17) 2πe σ2 β2 1 σ2 xσ2 x+σ2 z z , where Φ1 and Φ2 are the appropriate covariance matrices. The reason for (36) is that conditioned on xn−1 and yn, xn is independent of the other past samples of X and Y. Since X is drawn from i.i.d. Gaussian distribution with mean zero and variance σ2 x, the unconditional entropy of X is given by h (X) = 1 (cid:1) . Therefore, the DI from Y to X is 2 log(cid:0)2πeσ2 x I (Y → X)=h (X) − h (X(cid:107)Y)=1 2 log 1 + β2 1 σ2 σ2 z x . (38) (cid:16) (cid:17) C. Special cases 2 log x + σ2 z x + σ2 z h(xn, yn)=1 differential entropy of xn and yn is Consider the system in (17) with β1 = 1, β2 = 0. For this system, yn are i.i.d. Gaussian distributed with mean zero and 1 (cid:1). Therefore the differential entropy of YN variance(cid:0)σ2 is given by h(cid:0)YN(cid:1) = N (cid:1)(cid:1). Also the joint 2 log(cid:0)2πe(cid:0)σ2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:19) (cid:12)(cid:12)(cid:12)(cid:12)σ2 (cid:18) 2 log(cid:0)2πeσ2 (cid:1) . (39) =⇒ h(cid:0)YN(cid:107)XN(cid:1) = (cid:0)h (xn, yn) − N(cid:80) N(cid:80) (cid:1) . Therefore the directed information 2 log(cid:0)2πeσ2 h (xn)(cid:1) = N (cid:17) (cid:16) σ2 x x + σ2 z h (ynxn) = (cid:0)h(cid:0)YN(cid:1)−h(cid:0)YN(cid:107)XN(cid:1)(cid:1)=1 z from X to Y is given by 2 log The DI from Y to X can be similarly derived. I(X→Y)= lim N→∞ x σ2 x σ2 1+ σ2 x σ2 z (40) xσ2 z 2πe =1 n=1 n=1 . Now, consider the system in (17) with β1 = 0, β2 = 1. For this system, the DI from X to Y is computed by following the approach described above. Let us derive I (Y → X). The causal conditional entropy of XN given YN is given by h(cid:0)xnXn−1,YN(cid:1)= N(cid:80) h(xn)=h(cid:0)XN(cid:1) , (41) h(cid:0)XN(cid:107)YN(cid:1)= N(cid:80) n=1 n=1 since xn does not depend on the past samples of Y. Therefore, the DI from Y to X is zero, i.e., I (Y → X) = 0. REFERENCES [1] R. Malladi, G. Kalamangalam, N. Tandon, and B. Aazhang, "Identifying the epileptogenic zone using directed information," in Abstract presented at Soc. for Neuroscience (SfN), 2014. [2] R. Malladi, G. P. Kalamangalam, N. Tandon, and B. Aazhang, "Inferring causal connectivity in epileptogenic zone using directed information," in IEEE Int. Conf. on Acoust., Speech and Signal Process. (ICASSP), 2015, pp. 822 -- 826. [3] G. K. Bergey, M. J. Morrell, E. M. Mizrahi, A. Goldman, D. King- Stephens, D. Nair, S. Srinivasan, B. Jobst, R. E. Gross, D. C. Shields et al., "Long-term treatment with responsive brain stimulation in adults with refractory partial seizures," Neurology, vol. 84, pp. 810 -- 817, 2015. [4] F. Rosenow and H. Luders, "Presurgical evaluation of epilepsy," Brain, vol. 124, no. 9, pp. 1683 -- 1700, 2001. [5] E. Krook-Magnuson and I. Soltesz, "Beyond the hammer and the scalpel: selective circuit control for the epilepsies," Nature Neuroscience, vol. 18, no. 3, pp. 331 -- 338, 2015. [6] S. Sunderam, B. Gluckman, D. Reato, and M. Bikson, "Toward rational design of electrical stimulation strategies for epilepsy control," Epilepsy & Behavior, vol. 17, no. 1, pp. 6 -- 22, 2010. [7] K. J. Friston, "Functional and effective connectivity in neuroimaging: a synthesis," Human Brain Mapping, vol. 2, no. 1-2, 1994. [8] H. O. Luders, I. Najm, D. Nair, P. Widdess-Walsh, and W. Bingman, "The epileptogenic zone: general principles," Epileptic Disorders, vol. 8, 2006. [9] P. van Mierlo, M. Papadopoulou, E. Carrette, P. Boon, S. Vandenberghe, K. Vonck, and D. Marinazzo, "Functional brain connectivity from EEG in epilepsy: Seizure prediction and epileptogenic focus localization," Progress in neurobiology, vol. 121, pp. 19 -- 35, 2014. 15 [10] F. Pittau and S. Vulliemoz, "Functional brain networks in epilepsy: recent advances in noninvasive mapping," Current opinion in neurology, vol. 28, no. 4, pp. 338 -- 343, 2015. [11] S. L. Bressler and A. K. Seth, "Wiener -- Granger causality: a well established methodology," Neuroimage, vol. 58, pp. 323 -- 329, 2011. [12] C. W. Granger, "Investigating causal relations by econometric models and cross-spectral methods," Econometrica: J. of the Econometric Soc., 1969. [13] K. J. Blinowska, "Review of the methods of determination of directed connectivity from multichannel data," Medical & Biological Eng. & Computing, vol. 49, no. 5, 2011. [14] T. Schreiber, "Measuring information transfer," Physical Review Lett., vol. 85, no. 2, 2000. [15] E. Pereda, R. Q. Quiroga, and J. Bhattacharya, "Nonlinear multivari- ate analysis of neurophysiological signals," Progress in Neurobiology, vol. 77, no. 1, pp. 1 -- 37, 2005. [16] K. Lehnertz, "Epilepsy and nonlinear dynamics," J. of biological physics, vol. 34, no. 3-4, pp. 253 -- 266, 2008. [17] C. J. Quinn, T. P. Coleman, N. Kiyavash, and N. G. Hatsopoulos, "Esti- mating the directed information to infer causal relationships in ensemble neural spike train recordings," J. of Computational Neuroscience, vol. 30, no. 1, 2011. [18] K. So, A. C. Koralek, K. Ganguly, M. C. Gastpar, and J. M. Carmena, "Assessing functional connectivity of neural ensembles using directed information," J. of Neural Eng., vol. 9, no. 2, 2012. [19] N. Soltani and A. Goldsmith, "Directed information between connected leaky integrate-and-fire neurons," in IEEE Int. Symp. on Inform. Theory (ISIT), 2014, pp. 1291 -- 1295. [20] H. Marko, "The bidirectional communication theory -- a generalization of information theory," IEEE Trans. Commun., vol. 21, no. 12, 1973. [21] J. Massey, "Causality, feedback and directed information," in Int. Symp. on Inform. Theory Applications (ISITA), 1990. [22] G. Kramer, "Directed information for channels with feedback," Ph.D. dissertation, ETH Zurich, 1998. [23] J. Jiao, H. H. Permuter, L. Zhao, Y.-H. Kim, and T. Weissman, "Universal estimation of directed information," IEEE Trans. Inf. Theory, vol. 59, no. 10, pp. 6220 -- 6242, 2013. [24] Y. Liu, "Directed information for complex network analysis from multivariate time series," Ph.D. dissertation, Michigan State University, East Lansing, MI, USA, 2012. [25] Y. Liu and S. Aviyente, "The relationship between transfer entropy and directed information," in IEEE Statistical Signal Process. Workshop (SSP), Aug 2012, pp. 73 -- 76. [26] A. Rao, A. O. Hero, D. J. States, and J. D. Engel, "Inference of biolog- ically relevant gene influence networks using the directed information criterion," in IEEE Int. Conf. on Acoust., Speech and Signal Process. (ICASSP), vol. 2, 2006. [27] S. Tatikonda and S. Mitter, "The capacity of channels with feedback," IEEE Trans. Inf. Theory, vol. 55, no. 1, 2009. [28] H. H. Permuter, Y.-H. Kim, and T. Weissman, "Interpretations of di- rected information in portfolio theory, data compression, and hypothesis testing," IEEE Trans. Inf. Theory, vol. 57, no. 6, 2011. [29] P.-O. Amblard and O. J. Michel, "On directed information theory and Granger causality graphs," J. of Computational Neuroscience, vol. 30, no. 1, 2011. [30] T. Cover and J. Thomas, Elements of information theory. Wiley, 2006. [31] L. Barnett and A. K. Seth, "The MVGC multivariate Granger causality toolbox: a new approach to Granger-causal inference," J. of Neuro- science Methods, vol. 223, pp. 50 -- 68, 2014. [32] C. Diks and J. DeGoede, "A general nonparametric bootstrap test for Granger causality," in Global Analysis of Dynamical Systems, 2001. [33] D. N. Politis and J. P. Romano, "The stationary bootstrap," J. of the Amer. Statistical association, vol. 89, no. 428, pp. 1303 -- 1313, 1994. [34] C. Wilke, G. Worrell, and B. He, "Graph analysis of epileptogenic networks in human partial epilepsy," Epilepsia, vol. 52, no. 1, 2011. [35] S. Sabesan, L. B. Good, K. S. Tsakalis, A. Spanias, D. M. Treiman, and L. D. Iasemidis, "Information flow and application to epileptogenic focus localization from intracranial EEG," IEEE Trans. Neural Syst. Rehabil. Eng., vol. 17, no. 3, pp. 244 -- 253, 2009. [36] F. Panzica, G. Varotto, F. Rotondi, R. Spreafico, and S. Franceschetti, "Identification of the epileptogenic zone from stereo-EEG signals: a connectivity-graph theory approach," Frontiers in Neurology, 2013. [37] K. J. Blinowska and M. Kami´nski, "Multivariate signal analysis by para- metric models," Handbook of Time Series Analysis: Recent Theoretical Developments and Applications, p. 373, 2006. 16 [38] R. Malladi, G. P. Kalamangalam, and B. Aazhang, "Online bayesian change point detection algorithms for segmentation of epileptic activity," in Asilomar Conf. on Signals, Systems and Computers, 2013. [39] S. P. Meyn and R. L. Tweedie, Markov chains and stochastic stability. Cambridge University Press, 2009. [40] N.-Z. Shi and J. Tao, Statistical hypothesis testing: theory and methods. [41] S. M. Kay, Fundamentals of statistical signal processing : estimation World Scientific, 2008. theory. Prentice-Hall, 2010. [42] A. J. Izenman, "Review papers: recent developments in nonparametric density estimation," J. of the Amer. Statistical Assoc., vol. 86, no. 413, pp. 205 -- 224, 1991. [43] D. W. Scott, Multivariate density estimation: theory, practice, and visualization. John Wiley & Sons, 2015. [44] P. D. Grunwald, The minimum description length principle. MIT press, 2007. [45] A. R. Barron and T. M. Cover, "Minimum complexity density estima- tion," IEEE Trans. Inf. Theory, vol. 37, no. 4, 1991. [46] T. Duong et al., "ks: Kernel density estimation and kernel discriminant analysis for multivariate data in R," J. of Statistical Software, vol. 21, no. 7, pp. 1 -- 16, 2007. [47] D. Wied and R. Weissbach, "Consistency of the kernel density estimator: a survey," Statistical Papers, vol. 53, no. 1, pp. 1 -- 21, 2012. [48] G. Hu and R. O'Connell, "Analytical inversion of symmetric tridiagonal matrices," J. of Physics A: Math. and General, vol. 29, no. 7, 1996. [49] K. Ishiguro, N. Otsu, M. Lungarella, and Y. Kuniyoshi, "Comparison of nonlinear Granger causality extensions for low-dimensional systems," Physical Review E, vol. 77, no. 3, p. 036217, 2008. [50] C. Quinn, N. Kiyavash, and T. P. Coleman, "Directed information graphs," arXiv preprint arXiv:1204.2003, 2012. [51] P. Mierlo, E. Carrette, H. Hallez, R. Raedt, A. Meurs, S. Vandenberghe, D. Roost, P. Boon, S. Staelens, and K. Vonck, "Ictal-onset localization through connectivity analysis of intracranial EEG signals in patients with refractory epilepsy," Epilepsia, vol. 54, no. 8, pp. 1409 -- 1418, 2013. [52] C. Tonini, E. Beghi, A. T. Berg, G. Bogliun, L. Giordano, R. W. Newton, A. Tetto, E. Vitelli, D. Vitezic, and S. Wiebe, "Predictors of epilepsy surgery outcome: a meta-analysis," Epilepsy research, vol. 62, no. 1, pp. 75 -- 87, 2004. [53] Q. Wang, S. R. Kulkarni, and S. Verd´u, "Universal estimation of information measures for analog sources," Foundations and Trends in Communications and Information Theory, vol. 5, pp. 265 -- 353, 2009. [54] C. J. Stam, G. Nolte, and A. Daffertshofer, "Phase lag index: assessment of functional connectivity from multi channel EEG and MEG with diminished bias from common sources," Human brain mapping, vol. 28, no. 11, pp. 1178 -- 1193, 2007. [55] L. Faes and G. Nollo, Multivariate frequency domain analysis of causal INTECH Open Access time series. interactions in physiological Publisher, 2011. [56] C. P. Warren, S. Hu, M. Stead, B. H. Brinkmann, M. R. Bower, and G. A. Worrell, "Synchrony in normal and focal epileptic brain: the seizure onset zone is functionally disconnected," J. of neurophysiology, vol. 104, no. 6, pp. 3530 -- 3539, 2010. [57] S. P. Burns, S. Santaniello, R. B. Yaffe, C. C. Jouny, N. E. Crone, G. K. Bergey, W. S. Anderson, and S. V. Sarma, "Network dynamics of the brain and influence of the epileptic seizure onset zone," Proceedings of the National Academy of Sciences, vol. 111, no. 49, 2014. [58] R. Malladi, G. Kalamangalam, N. Tandon, and B. Aazhang, "Identifying seizure mechanisms from ECoG data using directed information," in Cosyne Abstracts, Salt Lake City, USA, 2016. Rakesh Malladi is a graduate student in Electrical and Computer Engineering Department at Rice University, Houston since fall 2011. He graduated with a dual degree (B.Tech & M.Tech) in Electrical Engineering from Indian Institute of Tech- nology Madras in 2011. He was a research intern at Cyberonics, Houston, USA dur- ing the summer of 2015, at Texas Instru- ments, Dallas, USA during the summer of 2013 and at IBM Research, Bangalore, India during the summer of 2010. His research interests spans topics in signal processing, machine learning, computational neuroscience and information theory. Giridhar Kalamangalam was born in Salem, India in 1965. He received the MBBS degree from the Jawahar- lal Institute in Pondicherry, India in 1989, and the MSc and DPhil degrees in applied mathematics from Oxford University, UK, in 1991 and 1995. Following postgraduate general medi- cal training (MRCP (UK)) he qualified as a neurologist (Glasgow, UK) with subspecialty expertise in epilepsy (Cleveland Clinic, USA). Since 2006, he has been with the University of Texas Health Science Center in Houston, TX, where he is currently Associate Professor of Neurology. His research interests are in the physiological dynamics and the neuroimaging of epilepsy and cognitive function. Nitin Tandon, MD, FAANS is a pro- fessor of neurosurgery at University of Texas Medical School at Houston. He received his medical degree at the Armed Forces Medical College in Pune, India, followed by a residency in neurosurgery at The University of Texas Health Science Center in San Antonio and a fellowship in epilepsy surgery at Cleveland Clinic. He has been in practice for thirteen years at Memorial Hermann TMC and on faculty at the UT Health Medical School. He has performed more than 3000 brain operations, with over 1200 for brain tumors and 600 for epilepsy. His lab performs multi- modality assessments of cognitive functions combining and correlating intracranial recordings, functional MRI, tractogra- phy and direct cortical stimulation. The thrust of his research is the development of optimal tools to characterize interactional brain processes across regions using intracranial recordings (www.tandonlab.org) Behnaam Aazhang received his B.S. (with highest honors), M.S., and Ph.D. degrees in Electrical and Computer En- gineering from University of Illinois at Urbana-Champaign in 1981, 1983, and 1986, respectively. In August 1985, he joined the faculty of Rice University, Houston, Texas, where he is now the J.S. Abercrombie Professor in the Department of Electrical and Computer Engineering Professor and Director of Center on Neuro-Engineering, a multi-university research center in Houston, Texas. He is a Fellow of IEEE and AAAS, and also a recipient of 2004 IEEE Communication Society's Stephen O. Rice best paper award for a paper with A. Sendonaris and E. Erkip. In addition, the authors received IEEE Communication Society's 2013 Advances in Communication Award for the same paper. He has been listed in the Thomson-ISI Highly Cited Researchers and has been keynote and plenary speaker of several conferences.
1202.4482
2
1202
2013-02-09T21:34:51
Metabolic cost as an organizing principle for cooperative learning
[ "q-bio.NC", "cs.LG", "nlin.AO" ]
This paper investigates how neurons can use metabolic cost to facilitate learning at a population level. Although decision-making by individual neurons has been extensively studied, questions regarding how neurons should behave to cooperate effectively remain largely unaddressed. Under assumptions that capture a few basic features of cortical neurons, we show that constraining reward maximization by metabolic cost aligns the information content of actions with their expected reward. Thus, metabolic cost provides a mechanism whereby neurons encode expected reward into their outputs. Further, aside from reducing energy expenditures, imposing a tight metabolic constraint also increases the accuracy of empirical estimates of rewards, increasing the robustness of distributed learning. Finally, we present two implementations of metabolically constrained learning that confirm our theoretical finding. These results suggest that metabolic cost may be an organizing principle underlying the neural code, and may also provide a useful guide to the design and analysis of other cooperating populations.
q-bio.NC
q-bio
Metabolic cost as an organizing principle for cooperative learning David Balduzzi1, Pedro A. Ortega2, Michel Besserve2 1Swiss Federal Institute for Technology, ETH Zurich 2Max Planck Institute for Intelligent Systems, Tubingen, Germany [email protected], {ortega, besserve}@tuebingen.mpg.de Abstract This paper investigates how neurons can use metabolic cost to facilitate learning at a population level. Although decision-making by individual neurons has been extensively studied, questions regarding how neurons should behave to cooperate effectively remain largely unaddressed. Un- der assumptions that capture a few basic features of cortical neurons, we show that constraining reward maximization by metabolic cost aligns the information content of actions with their expected reward. Thus, metabolic cost provides a mechanism whereby neurons encode expected reward into their outputs. Further, aside from reducing energy expenditures, impos- ing a tight metabolic constraint also increases the accuracy of empirical estimates of rewards, increasing the robustness of distributed learning. Fi- nally, we present two implementations of metabolically constrained learning that confirm our theoretical finding. These results suggest that metabolic cost may be an organizing principle underlying the neural code, and may also provide a useful guide to the design and analysis of other cooperating populations. 1 Introduction Rational decision making is typically formalized as optimizing a reward function [19,22,26]. This paper investigates neuronal learning from an optimization perspective. We assume that both the brain as a whole and individual neurons are rational decision makers optimizing reward functions of some kind. Since the brain learns from finite samples, it is exposed to a tradeoff between over- and under-fitting: increasing a model's capacity can improve its fit on training data, but poten- tially worsens performance on future samples [29]. Remarkably, however, the human brain effortlessly handles a wide-range of complex pattern recognition tasks suggesting it both has a large capacity and, paradoxically, also generalizes extremely well. On the basis of these conflicting observations, it has been argued that useful biases in the form of "generic mechanisms for representation" must be hardwired into cortex [10]. Our goal in this paper is to propose a bias that is both useful and biologically plausible. Let us outline the problem. Neurons learn inductively. They can generalize from finite samples and encode estimates of future outcomes (for example, rewards) into their spike- trains [12]. Learning-theoretical results imply that generalizing successfully from small sam- ples requires strong biases [29] or, in other words, specialization. Thus, at any given time some neurons' specialties are more relevant than others. Since most of the data neurons receive are other neurons' outputs, it is essential that they indicate which of their outputs 1 encode high quality estimates. Downstream neurons should then be biased to specialize on these outputs, thereby reducing the effective search space that neurons explore. The question we ask is: How can a single neuron near-maximize its expected reward and simultaneously help downstream neurons do the same? As a partial answer, we propose the following organizing principle. Neurons should consis- tently label outputs as useful and not useful. Specifically, the optimizations performed by neurons should be designed so that spikes are useful and silences are not. For example, it is well known that, on average, the more spikes a neuron receives, the more it learns by modi- fying its synapses. Although this fact is often taken for granted, it requires explanation. We suggest that spikes drive learning because they are useful. By useful, we mean an output that (i) predicts high reward with (ii) tight confidence intervals. Distinguishing useful from irrelevant outputs is helpful to downstream neurons since it reduces the size of the search space they are confronted with. If spikes reliably predict future reward, then neurons should be biased towards learning from spikes. In fact, this fits well with experimental evidence [8, 9, 20]. Moreover, consistently biasing learning toward outputs labeled as useful (i.e. spikes) also provides a principled way to reduce capacity, and thus improve generalization guarantees, without sacrificing empirical performance. The scope of this paper is limited to showing how neurons may systematically distinguish useful from irrelevant outputs. Fleshing out the implications for learning at the population level is deferred to future work. Furthermore, we only consider excitatory connections in this paper; the role of inhibition is also deferred to future work. Overview. We explore the consequences of a few basic assumptions regarding spikes, metabolic cost and neuronal rewards, see §2. We consider a minimal model of biological neural network from which we extract guiding principles that may apply, at least approx- imately, to more complex, biologically realistic models. The most important assumptions are that neurons aim to maximize reward after spiking, and that neurons have to operate within a fixed metabolic budget that constrains how often they can spike in a given time interval. It turns out that these seemingly innocuous assumptions are the key to distinguishing useful from irrelevant outputs. Our first result is that constrained reward maximization causes neurons to maximize the information they encode into their spikes, see §3. Moreover, if spikes are sufficiently rare, it turns out that spikes dominate the information communicated by neurons, which has interesting implications for credit assignment in cortex, see §3 and [5]. If neurons attempt to maximize empirical reward after spiking, it follows that neurons encode reward estimates into spikes. When a neuron produces a spike, it thus signals to downstream neurons that it expects a positive neuromodulatory signal such as dopamine. This raises questions concerning the quality of the reward estimates encoded in spikes, see §4. We show that the more information neurons encode into spikes, the tighter the guarantees tying empirical reward to expected reward. We conclude by describing some implications of our results for cooperative optimization. Theorems are proved in the appendices. Related work. Neuronal plasticity and its implications for the neural code have been intensively studied for many years. The work closest in spirit to this paper is Seung's "hedonistic" synapses, which seek to increase average reward [24]. A second related line of research applies the information bottleneck method – an alternate constraint to the one considered here – to neuronal learning [7, 27]. An information-theoretic perspective on synaptic homeostasis that complements the results in this paper is [5]. The consequences of constrained optimization for spike-timing dependent plasticity (STDP) are presented in [2]. A practical implementation of regularized STDP, inspired by the ideas presented here, can be found in [16]. Acknowledgements. We thank Yevgeny Seldin and Giulio Tononi for useful discussions. 2 2 A minimal model The mammalian cortex contains between 107 and 1011 neurons (depending on the species) that are guided by neuromodulators signaling pleasure, pain and other globally salient events. Neurons communicate with each other via spiketrains – sequences of silences and spikes they receive from and transmit to 103 to 104 other neurons through connections called synapses. Neurons learn by increasing and decreasing the efficacy of their synapses according to the timing of pre-synaptic (input) and post-synaptic (output) spikes, as well as the presence or absence of neuromodulators [8, 9, 15, 20, 25]. Model neurons. The cortex can be modeled as a population of K neurons. Neuron nk follows policy πk, where πk(as) is a Markov matrix specifying the probability the neuron picks action, or output, a ∈ A = {0, 1} upon encountering situation s ∈ S. A situation, or input, is a vector s = (a1, . . . , aK) whose entries are the actions of all the neurons in the brain at the previous time step. Since each neuron is only exposed to a small fraction of the brain, it ignores most entries in the vector. Thus, the policy of neuron k is πk(as) = πk(aΣk) for some subset Σk ⊂ {a1, . . . , aK}. Let P (S) denote the prior over situations. We will assume the prior is i.i.d. when providing guarantees on estimates, see Remark 4 for a brief discussion. Neurons are exposed to a global neuromodulatory signal ν ∈ N that signals the performance of the population as a whole. Neuromodulators are drawn with probability P (νs). We make the following assumptions. Assumption 1 (reward maximization). Neurons maximize a reward function that depends on neuromodulatory signals, input spikes and output spikes: Rk(si, ai, νi), (1) (cid:98)πk = arg max π∈M N(cid:88) i=1 where ai is the output chosen by π in response to input si. The set M is the set of possible neuronal mechanisms, a subset of the set Markov matrices on S × A. Two examples of M that are relevant to our discussion are discrete threshold neurons M =(cid:8)H((cid:104)w, s(cid:105) − ϑ)(cid:12)(cid:12) w ∈ RK(cid:9) where H(•) is the Heaviside function, see [2], and the full set of Markov matrices, see discussion of Q-learning below. Note that, since different neurons are exposed to different subsets of the total brain activity, they have different reward structures, even for the same neuromodulators: in general Rk (cid:54)= Rj for j (cid:54)= k since Σk (cid:54)= Σj. Since we focus on the behavior of a single neuron, we will often drop the subscript k from the notation below. Assumption 2 (spikes gate rewards). The reward function is gated by synaptic outputs: R(s, a, ν) = R(s, ν) · Ia=1, where Ia=1 = is the indicator function. (2) (cid:26)1 a = 1 0 else Neurophysiological evidence suggests that neurons only potentiate or depotentiate their synapses shortly before or after producing spikes [8,9,15,25]. This encourages specialization: neurons search for a small set of inputs that reliably predict future reward signals – other inputs are ignored. For simplicity we assume neurons have two outputs, spikes and silence. We use notations a0 or a = 0 for silence and a1 or a = 1 for spikes. 3 Assumption 3 (metabolic budget). Neurons have a fixed metabolic budget that determines the maximum frequency of spiking over some (unspecified) time period: where π(a1) =(cid:80) s p(s) · π(a1s) is the spiking frequency under policy π. π(a1) ≤ ρ (3) Although a soft constraint is more biologically plausible, imposing a hard constraint simpli- fies the exposition without significantly altering the conclusions. The main effect of softening the constraints is to allow the information carried by spikes and the capacity of neurons to vary, thereby softening Theorems 2 and 3 below. Since neurons only modify their synapses when they spike, it follows that neurons that spike very infrequently learn very little if at all. Thus, we expect that there are mechanisms in place ensuring that neurons not only stay within their metabolic budget, but also that they come close to using all of it. Spikes and silences are not abstract, interchangeable symbols. Spiking and responding to spikes carries a much higher metabolic cost than not spiking [13]. This cost is significant since the nervous system consumes a disproportionate share of an organism's total energy budget [1]. It has been hypothesized that a function of sleep is to homeostatically regulate synaptic strengths, so as to control metabolic expenditures associated with action potentials, see [11, 14, 28, 30]. Unlike "agents in the wild", individual neurons have negligible impact on what happens next: a single neuron has little effect on the neurons it targets – since each receives inputs from thousands of other neurons. The inability of individual neurons to manipulate their environment simplifies the optimization problems they face by stripping out the recursive Bellman aspect [6]: Assumption 4 (disempowerment). An individual neuron has no immediate influence on its environment (cid:12)(cid:12)(cid:12) s(t), ak(t) (cid:17) (cid:16) (cid:12)(cid:12)(cid:12) s(t) (cid:17) p s(t + 1, . . .) = p s(t + 1, . . .) , (4) (cid:16) where t refers to time. I.e. conditioning on the neuron's output makes no difference to the distribution on subsequent actions by the rest of the brain. The future situations a neuron encounters are essentially unaffected by its output. This assumption fails at the population level – populations of neurons necessarily affect the organism's actions. Nevertheless Eq. (4) is a reasonable assumption at the individual neuron level since, for example, destroying a single neuron makes essentially no difference to brain function. 3 Encoding useful information in spikes This section considers the implications of our assumptions for the information content of spikes. Mutual information provides a formal method for quantifying the information con- tent of a channel. We present a related measure, effective information, that measures the information content of a single output in terms of how much that output reduces prior uncertainty regarding the set of inputs. Importantly, we show that under the above as- sumptions, spikes not only reduce uncertainty, but also signal that the neuron received an input that was historically followed by high rewards. The information encoded in spikes is thus indicative of expected future rewards. Information. It is useful to consider neurons as communication channels mapping situ- ations to actions. The average information communicated by a neuron is then the mutual information Iπ(S,A). However, since we are interested in the information communicated by specific actions (in particular, spikes), we introduce effective information1 [3, 4] which quantifies the information that a single output encodes about the input. 1We extend the definition of effective information in [3] to allow arbitrary priors instead of restricting to the uniform distribution. 4 Definition 1 (effective information). Given a neuron with policy π(as) and prior P (S) on situations, the effective information generated by action a ∈ A is ei(π, a) := D(cid:2)π(Sa)(cid:13)(cid:13) P (S)(cid:3) where π(sa) := π(as) π(a) · P (s) is computed via Bayes' rule and D[•(cid:107)•] is the Kullback-Leibler divergence D[p(cid:107)q] (5) := (cid:80) pi log pi qi . An interesting special case is when the prior on situations is the uniform distribution and the policy is deterministic. It follows that ei(π, a) = − log π−1(a) S , (6) where • denotes cardinality and (since the policy is deterministic), π is a function π : S → A. In Eq (6), effective information quantifies the selectivity of an output: the fraction of inputs causing the policy to output a. The smaller the fraction, or alternatively the more sensitive output a is to perturbations in the input, the higher effective information [5]. Remark 1. Suppose we have model PM(dh) that specifies the probability of observing data given a hypothesis. Further suppose we have prior distribution P (h) on hypotheses. If we observe data d, how much have we learned about the hypotheses? The Bayesian information gain is (7) If we consider a neuron's policy as a model, with inputs as hypotheses and outputs as evidence, then effective information quantifies the Bayesian information gained about the inputs given an output. Remark 2. The expectation of ei is mutual information: E (cid:2)ei(π, a)(cid:3) = Iπ(S,A). π(a) D(cid:2)PM(Hd)(cid:13)(cid:13) P (H)(cid:3). (cid:98)R(s, a) := 1 N (cid:88) Information aligns with rewards (theory). We show that neurons implementing con- strained reward maximization from Assumption 1 also maximize the effective information of their spikes ei(π, a1), that we will call information per spike. Definition 2 (empirical reward). Given a finite sample of situations, actions and neuromodulators (si, ai, νi)N pirical reward observed after performing action a in situation s be i=1, let the em- R(si, ai, νi). {isi=s,ai=a} We also introduce the empirical reward after spiking, 1 N pirical reward per spike (cid:1) and the em- (cid:80){iai=1} R(cid:0)si, a1, νi Assume that situations yield different empirical rewards, i.e. (cid:98)R(s, a1) (cid:54)= (cid:98)R(s(cid:48), a1) for all s, s(cid:48) ∈ S. Recall that by Assumptions 1 and 3 an optimal policy (cid:98)π satisfies Theorem 1 (maximizing reward/spike maximizes information/spike). (cid:80){iai=1} R(cid:0)si, a1, νi N(cid:88) {iai=1} (cid:1). 1 (cid:98)π = arg max {π∈Mπ(a1)≤ρ} R(si, ai, νi), i=1 If we also apply constraint π(a1) ≥ ρ, i.e. ρ is both upper and lower bound, then the optimal policy (cid:98)π maximizes information per spike. More precisely, the optimal policy (cid:98)π satisfies Remark 3. The optimal policy will satisfy(cid:98)π(a1) = ρ if there are enough inputs s satisfying ei((cid:98)π, a1) ≥ ei(π, a1) for all π such that π(a1) = ρ. Eν[R(s, ν, a1)] > 0. In other words, if there are enough situations where spiking, on average, is followed by positive reward. Imposing the lower bound in the theorem means we compare the optimal policy with alternate policies that spike with the same frequency. Thus, the optimal policy necessarily maximizes both the empirical reward after spiking and the effective information encoded in spikes. We illustrate this result with simulations below. 5 Information aligns with rewards (experiments). A learning algorithm implementing constrained reward maximization is a modification of Q-learning [31]: Example 1 (Metabolically constrained Q-learning). If a neuron chooses action a in situation s and subsequently receives neuromodulator ν, then it updates the Q-matrix by Q(s, a) ← Q(s, a) + α ·(cid:104)(cid:98)R(s, a, ν) − Q(s, a) (cid:105) , where α controls the rate. After updating Q, the neuron constructs new policy π(as) = Mn(cid:16) eQ(s,a)(cid:17) . Operation M(•) s∈S π(as)P (s) = P (a) for all a, and then by Z(s) chosen such that (cid:80) (cid:80) first by Z(a) chosen such that a∈A π(as) = 1 for all s. Setting n = 3 yields a policy that approximately implements the metabolic con- straint. the policy twice: renormalizes Figure 1 shows how effective information and empirical reward after spiking covary as neu- rons Q-learn. We initialized 5000 neurons randomly and applied metabolic constraints ρ ∈ {0.1, 0.3, 0.5}. Rewards are drawn randomly. As the neurons adapt, their policies be- come both more deterministic and more likely to spike in situations yielding higher rewards, so as neurons adapt they both encode more information into their spikes and predict higher rewards after spiking. The tighter the metabolic constraint (i.e. the lower ρ), the higher the empirical reward after spiking. Thus, the information encoded in spikes provides a reliable guide to the empirical reward after spiking. Spikes dominate information content. Theorem 1 shows information per spike is max- imized by constrained reward maximization. Theorem 2 below consider how much of the total information communicated by a neuron is carried by spikes. Theorem 2 (spikes carry essentially all information). Suppose a neuron has two actions (silence a0 and spike a1) and produces spikes infrequently: π(a1) (cid:28) 1. Then the total information communicated by the neuron is approximately the information it communicates using spikes alone: Iπ(S;A) = π(a1) · ei(π, a1) + O(cid:0)π(a1)2(cid:1) . (8) If the metabolic constraint is tight, meaning ρ and so π(a1) are small, then for the optimal policy encodes a lot of information into spikes. In this setting, Theorem 2 implies that the information communicated by a neuron is (up to first order) carried by spikes alone. Eq (6) and Theorem 2 together have interesting implications for credit assignment, see [5] for details. In particular, if spikes carry most of the information in cortex, then spiking neu- rons and synapses should reinforced in response to positive global signals such as dopamine, and conversely for negative global signals. Neurons and synapses that are silent contribute little to the information generated by cortex, and so should be neither potentiated nor depo- tentiated. This fits neurophysiological evidence suggesting that spikes play a distinguished role in synaptic potentiation and depotentiation [8, 9, 15, 20]. pirical rewards(cid:8)(cid:98)R(s, a1)(cid:12)(cid:12)s ∈ S(cid:9). The optimal policy is constructed as follows. First, rank Encoding reward estimates in spikes. Finally, we briefly illustrate the effect of the metabolic constraint ρ on the empirical reward per spike. Suppose we have sampled em- states by their empirical reward. Let Sρ denote the states in the top ρth percentile. Define (cid:98)π(a1s) = s ∈ Sρ else. (cid:26)1 0 The optimal policy can be visualized as moving a window of fixed size and variable shape over the input space, such that the underlying configuration maximizes reward. 6 Figure 2 shows an example. Situations are ranked according to their empirical reward. The metabolic constraint is set at ρ = 15%, so the optimal policy spikes for the 15% of situations with highest reward. The policy picking these situations results in a reward per spike of 0.89. Situations which do not cause spikes receive no reward ; if spikes and silences would be exchanged, this reward would be .21 (we call it reward after silence). We make two observations. First, tightening the metabolic constraint, so that the policy spikes for (cid:28) 15% of situations, increases the reward per spike. Second, the variance in reward per spike is much lower than after not spiking, and typically decreases with ρ. As a general rule of thumb, tightening the metabolic constraint by decreasing ρ both in- creases the empirical reward per spike and reduces the variance in empirical reward per spike. 4 Guarantees on reward estimates In a dynamically changing environment like the cortex, it is important that neurons reliably represent high reward. This section shows that the reliability of spikes depends on how much information is encoded in them. Theoretical guarantees. We say that a neuron's spikes reliably represent reward when there is a low variability in the empirical mean reward when the neuron spikes. For a given policy π. The expected and empirical rewards per spike are Rπ := E(cid:104) (cid:98)Rπ := R(S, a1, N ) (cid:88) s,n (cid:105) (cid:88) := 1 T1 {(st,νt)π(st)=1} π(sa1) · P (νs) · R(s, a1, ν) and R(cid:0)st, a1, νt (cid:1) respectively, where S is the total number of possible situations. Let πu(a) =(cid:80) where T1 counts spikes produced by the neuron during [1, T ]. We measure sample size in terms of spikes (rather than spikes and silences) because spikes are metabolically expensive, so only spikes count towards the cost of collecting a finite sample. Let us introduce notation for computations with respect to the uniform prior U (s) = 1S , s π(as)· U (s), πu(sa) = π(as) U (s) πu(a) . Note that eiu recovers the original notion of effective information in [3], which Definition 1 generalizes. π(a) , and eiu(π, a) =(cid:80) s πu(sa) log πu(sa) The following theorem is proved using a version of Occam's razor [23]. Theorem 3 (error bound for empirical reward). Suppose the neuron chooses a deterministic policy π under the constraint that it spikes for a fixed fraction of situations: πu(a1) = const. Further, suppose that situations are sampled i.i.d. Without loss of generality,2 assume that rewards lie in [0, b]. Then with probability at least 1 − δ, (cid:115) (cid:12)(cid:12)(cid:12)Rπ − (cid:98)Rπ (cid:12)(cid:12)(cid:12) ≤ b · Guarantees improve as eiu increases since x+1 S · eiu(π, a1) + 1 2T1 · eeiu(π,a1) + ex decreases as x increases. log 2 δ 2T1 (9) Note that ei and eiu covary since increasing the number of situations where a neuron spikes decreases both ei and eiu; similarly, decreasing the number of situations where a neuron spikes increases both ei and eiu. Thus, tightening the metabolic constraint in Assumption 3, by choosing low ρ, yields policies that have better guarantees on their reward estimates. 2Since only relative rewards affect the choice of optimal policy, it follows that negative rewards can be stripped out of the optimization problem by introducing an additive constant. 7 Remark 4. The assumption that inputs are i.i.d. is not realistic for cortical neurons. We make two remarks. First, if rewards are only non-zero in the presence of neuromodulatory signals, then the assumption states that situations directly preceding neuromodulator release are i.i.d, which is more reasonable. Second, similar results have recently been obtained in non-i.i.d. scenarios using more sophisticated PAC-Bayes methods [21] – and these may be applicable to our setting. Guarantees in practice. Figure 3 plots information encoded in spikes (x-axis) against the difference between the normalized empirical and expected reward (y-axis). Rewards were drawn randomly and the expected and empirical error for 16,000 deterministic policies with S = 50, P (S) uniform and T1 = 20 were computed. Policies were sampled randomly with k, the number of situations causing the policy to spike, varying uniformly across [1, 25]. The figure shows that both normalized error and the standard error of the error decrease as ei increases. Figure 3 confirms that the bound in Theorem 3 is a reasonable guide to performance in practice. Thus, the more information encoded in spikes, the better a neuron's empirical estimate of its expected reward per spike. The metabolic constraint in Assumption 3 controls the quality of a neuron's empirical estimates of its expected reward. 5 Discussion The space of possible policies that the cortex as a whole could implement is vast: it consists in choosing synaptic weights of millions or billions of neurons each receiving inputs from thousands or tens of thousands of other neurons. However, the space of policies that makes sense biologically is probably much smaller due to biological as well as learning-theoretic considerations. This paper has shown that metabolically constrained reward maximization provides a biologically plausible way for neurons to distinguish useful outputs from those that are not. It follows from our assumptions that spikes: • are responsible for most of the information communicated by neurons; • signal when, based on empirical estimates, neurons predict high reward; and • come equipped with performance guarantees that increase as the metabolic con- straint is tightened, and so more information is encoded in spikes. This suggests that neurons should privilege spikes during learning; and indeed there is a large body of experimental evidence that this is exactly what occurs [8, 9, 15, 25]: synaptic plasticity is triggered by pre- and post-synaptic spikes, with the decision to potentiate or depotentiate depending on their precise timing. An important unanswered question is setting the metabolic constraint ρ. Clearly, if ρ is too low, then neurons will barely fire at all, which is not desirable. Conversely, if ρ is too high then information and empirical reward encoded in spikes, as well as the quality of the empirical estimates, all degrade, which is also to be avoided. In this paper, we have simply highlighted the metabolic constraint ρ as an important lever that neurons may actively manipulate. Finding an optimal value or range of values for ρ is outside the scope of this paper. Biasing neuronal mechanisms. Q-learning is not a practical learning rule; it simply constructs a giant lookup table. If we think of the set of inputs causing a neuron to fire as a window of varying shape, then the metabolic constraint fixes the size of the window and Q-learning places no other constraints on its shape. However, if upstream neurons systematically encode information and reward in spikes, then it makes sense to bias shape of downstream neuronal "firing windows" to take the asymmetry between spikes and silences into account. This is exactly what we occurs in cortex. The vast majority of synapses are excitatory: the more input spikes neurons receive, the more likely they are to spike themselves. 8 Cortical neurons are biased toward firing for more spikes, which makes sense if spikes are reliable predictors of future reward. Thus, neuronal "firing windows" are shaped such that if a firing pattern causes a neuron to spike, then so does any firing pattern containing strictly more spikes. Moreover, sufficiently many pre-synaptic spikes will (essentially) always cause a neuron to fire. Thus, neurons aggregate evidence for high reward (spikes from upstream neurons), and modify their synaptic strengths to maximize their empirical reward. Fine-tuning is necessary since no two neurons have exactly the same connectivity, and therefore no two neurons use the same data to predict global neuromodulatory signals. Inhibitory (GABA) synapses do not fit this picture. Extending our framework for plasticity to include both types of neurons is challenging. However, we conjecture this incompatibility can be overcome, with inhibition playing a complementary role; possibly centered on im- posing sparse activity in the brain [17, 18] and selecting competing neural assemblies and structures. Global versus overlapping local optimizations. The results in this paper depend on specific assumptions and model choices. A particularly important assumption is that all neurons attempt to maximize the same global neuromodulatory reward signal. In this scenario, since neurons differ in their connectivity, they have access to different subsets of brain activity representing different environemental features and events, and therefore specialize on different sources of reward. The reality is more complicated with multiple overlapping neuromodulatory signals includ- ing dopamine, noradrenaline, acetylcholine and others. Moreover, neurons involved in, say, early visual processing may not require neuromodulatory guidance; rather, they may search for stable invariants over short time frames (hundreds of milliseconds). There is likely a diverse array of reward functions implemented across cortex. It is thus unclear whether the brain can be accurately described as optimizing a single well- defined reward function. Nevertheless, until decisively shown to be false, we believe the optimization perspective to be a fruitful working hypothesis. Even if it does not apply to the brain as a whole, it may nevertheless provide insight into how populations of neurons in specific brain areas converge on useful behaviors. A spiking currency. Finally, it is interesting to speculate on an analogy between spikes and paper currency. Money plays many overlapping roles in an economy, including: (i) focusing attention; (ii) stimulating activity; and (iii) providing a quantitative lingua franca for tracking revenues and expenditures. Note that, as for the brain, it is unclear whether an economy as a whole can be reduced to optimizing a single well-defined function. Spikes may play similar roles in cortex to those of paper currency in an economy. Spikes focus attention: STDP and other proposed learning rules are particularly sensitive to spikes and spike timing. Spikes stimulate activity: input spikes cause output spikes. Finally, spikes leave trails of (Calcium) traces that are used to reinforce and discourage neuronal behaviors in response to neuromodulatory signals. Neither money nor spikes are intrinsically valuable. Currency can be devalued by inflation. Similarly, the information content and guarantees associated with spikes can be eroded by overpotentiating synapses which reduces their selectivity (potentially leading to epileptic seizures in extreme cases). Regulating the information content of spikes is therefore essential. Assumption 3 provides a simple constraint that can be approximately imposed by regulating synaptic weights. Indeed, there is evidence that one of the functions of sleep is precisely this [11, 14, 28, 30]. Spikes with high information content are valuable because they come with strong guarantees on their estimates. They are therefore worth paying attention to, worth responding to, worth keeping track of, and worth learning from. 9 References [1] Attwell, D. and Laughlin, S. B., An energy budget for signaling in the grey matter of the brain, J Cereb Blood Flow Metab 21 (2001) 1133–45. [2] Balduzzi, D. and Besserve, M., Towards a learning-theoretic analysis of spike-timing de- pendent plasticity, Advances in Neural Information Processing Systems (NIPS) (2012). [3] Balduzzi, D. and Tononi, G., Integrated Information in Discrete Dynamical Systems: Motivation and Theoretical Framework., PLoS Comput Biol 4 (2008) e1000091. [4] Balduzzi, D. and Tononi, G., Qualia: the geometry of integrated information, PLoS Comput Biol 5 (2009) e1000462. [5] Balduzzi, D. and Tononi, G., What can neurons do for their brain? Communicate selectivity with spikes, Theory in Biosciences (2012). [6] Bellman, R., Dynamic Programming (Princeton University Press, 1957). [7] Buesing, L. and Maass, W., Simplified rules and theoretical analysis for information bottleneck optimization and PCA with spiking neurons, in Adv in Neural Information Processing Systems (NIPS) (2007). [8] Dan, Y. and Poo, M.-M., Spike timing-dependent plasticity of neural circuits., Neuron 44 (2004) 23–30. [9] Dan, Y. and Poo, M.-M., Spike timing-dependent plasticity: from synapse to percep- tion., Physiol Rev 86 (2006) 1033–1048. [10] Geman, S., Bienenstock, E., and Doursat, R., Neural Networks and the Bias/Variance Dilemma, Neural Comp 4 (1992) 1–58. [11] Gilestro, G. F., Tononi, G., and Cirelli, C., Widespread changes in synaptic markers as a function of sleep and wakefulness in Drosophila, Science 324 (2009) 109–12. [12] Gottfried, J. A., O'Doherty, J., and Dolan, R. J., Encoding predictive reward value in human amygdala and orbitofrontal cortex, Science 301 (2003) 1104–7. [13] Hasenstaub, A., Otte, S., Callaway, E., and Sejnowski, T. J., Metabolic cost as a unifying principle governing neuronal biophysics, Proc Natl Acad Sci U S A 107 (2010) 12329–34. [14] Maret, S., Faraguna, U., Nelson, A. B., Cirelli, C., and Tononi, G., Sleep and waking modulate spine turnover in the adolescent mouse cortex., Nat Neurosci 14 (2011) 1418– 1420. [15] Markram, H., Lubke, J., Frotscher, M., and Sakmann, B., Regulation of synaptic effi- cacy by coincidence of postsynaptic aps and epsps, Science 275 (1997) 213–5. [16] Nere, A., Olcese, U., Balduzzi, D., and Tononi, G., A neuromorphic architecture for object recognition and motion anticipation using burst-STDP, PLoS One 7 (2012) e36958. [17] Olshausen, B. A. and Field, D. J., Sparse coding with an overcomplete basis set: a strategy employed by v1?, Vision Res 37 (1997) 3311–25. [18] Olshausen, B. A. and Field, D. J., Sparse coding of sensory inputs, Curr Opin Neurobiol 14 (2004) 481–7. [19] Ortega, P. A. and Braun, D. A., Information, utility and bounded rationality, The fourth conference on artificial general intelligence (2011) 269–274. [20] Pawlak, V., Wickens, J. R., Kirkwood, A., and Kerr, J. N. D., Timing is not everything: neuromodulation opens the STDP gate, Front. Syn. Neurosci 2 (2010). [21] Rubin, J., Shamir, O., and Tishby, N., Trading Value and Information in MDPs, in Decision Making with Imperfect Decision Makers, eds. Guy, T. V., K´arn´y, M., and Wolpert, D. (Springer, 2011). [22] Russell, S. and Norvig, P., Artificial Intelligence: A Modern Approach, 3rd edn. (Pren- tice Hall, 2009). [23] Seldin, Y. and Tishby, N., Multi-Classification by Categorical Features via Clustering, in Proceedings of the 25th International Conference on Machine Learning (2008). 10 [24] Seung, H. S., Learning in Spiking Neural Networks by Reinforcement of Stochastic Synaptic Transmission, Neuron 40 (2003). [25] Song, S., Miller, K. D., and Abbott, L. F., Competitive Hebbian learning through spike-timing-dependent synaptic plasticity, Nature Neuroscience 3 (2000). [26] Sutton, R. S. and Barto, A. G., Reinforcement Learning: An Introduction (MIT Press, 1998). [27] Tishby, N., Pereira, F., and Bialek, W., The information bottleneck method, in Proc. of the 37-th Annual Allerton Conference on Communication, Control and Computing, eds. Hajek, B. and Sreenivas, R. (1999). [28] Tononi, G. and Cirelli, C., Sleep and synaptic homeostasis: a hypothesis, Brain Res. Bull. 62 (2003) 143–150. [29] Vapnik, V., Estimation of Dependencies Based on Empirical Data (Springer, 1982). [30] Vyazovskiy, V. V., Olcese, U., Lazimy, Y., Faraguna, U., Esser, S. K., Williams, J. C., Cirelli, C., and Tononi, G., Cortical firing and sleep homeostasis, Neuron 63 (2009) 865–78. [31] Watkins, C. and Dayan, P., Q-learning, Machine Learning 8 (1992) 279–292. A.1 Proof of Theorem 1 Proof. First we show that deterministic policies maximize the information encoded in spikes, then we show that deterministic policies maximize reward per spike. Deterministic policies maximize effective information subject to π(a1) ≤ ρ. Observe that (cid:88) (cid:88) s∈S s∈S π(sa) log π(sa) log − log π(a) (cid:124) (cid:123)(cid:122) π(sa) P (s) π(as) π(a) (cid:125) log frequency of output a ei(π, a) = = = (cid:88) (cid:124) s∈S + π(sa) log π(as) . (cid:125) (cid:123)(cid:122) stochasticity The log-frequency, or surprise, term is nonnegative and the stochasticity term is non- positive. It is easy to see that the stochasticity term is maximized at 0 if and only if output a is chosen deterministically – i.e. π(as) is either 0 or 1 for all s ∈ S. Deterministic policies maximize reward per spike. Suppose there are N situations ordered ministic policy spiking only for the ρ · N policies with highest empirical reward maximizes empirical reward after spiking. according to their empirical reward, so that (cid:98)R(s1) < ··· < (cid:98)R(sN ). It is clear that a deter- A.2 Proof of Theorem 2 Proof. Observe that P (s) − π(sa1) · π(a1) P (sa0) = Thus to first order in π(a1), π(sa0) is of the form p + δp where (cid:82) δp = 0. We can then = P (s) + π(a1)(cid:0)P (s) − π(sa1)(cid:1) + O(cid:0)π(a1)2(cid:1). 1 − π(a1) compute D(cid:2)p + δp (cid:13)(cid:13) p(cid:3) = (cid:90) (cid:90) (p + δp) log2 = α (p + δp) δp p p + δp (cid:18) p 1 − δp 2p (cid:19) + O(cid:0)(δp)3(cid:1) = α (cid:90) (δp)2 2p + O(cid:0)(δp)3(cid:1) , 11 ln 2 . where α = 1 Substituting p = P (s) and δp = π(a1)(cid:0)P (s) − π(sa1)(cid:1) gives D(cid:2)π(Sa0)(cid:13)(cid:13) P (S)(cid:3) = D(cid:2)p + δp(cid:13)(cid:13) p(cid:3) + O(cid:0)π(a1)2(cid:1) (cid:90) (cid:0)P (s) − π(sa1)(cid:1) Thus I(S; A) = π(a1)D(cid:2)π(Sa1)(cid:13)(cid:13) P (S)(cid:3) + O(cid:0)π(a1)2(cid:1). 2P (s) = α · π(a1)2 + O(cid:0)π(a1)2(cid:1). A.3 Proof of Theorem 3 Occam's razor can be paraphrased to say that the simplest hypothesis should be preferred. Suppose we have a setof hypotheses H with prior distribution P (h) on H. Let − log P (h) denote the complexity of hypothesis h. Let L : X × H → [0, b] be a loss function. Then Theorem 4 (Occam's razor). For any data generating distribution on X and any prior distribution P (h) over H, with a probability greater than 1− δ over drawing an i.i.d. sample from X of size T , for all h ∈ H: (cid:12)(cid:12)(cid:12)L(h) −(cid:98)L(h) (cid:12)(cid:12)(cid:12) ≤ b · (cid:115)− log P (h) + log 2 δ . 2T Proof. See [23]. Let H =(cid:8)π : S → A(cid:9) denote the set of deterministic policies, where A = {a0, a1} = {0, 1}. Define loss function L : (cid:16)S × N(cid:17) × H −→ R : (s, ν) × π (cid:55)→ R(s, π(s), ν). Further, set probability distribution P (s, ν) = P (s) · P (νs) on S × N . Theorem 4 holds for any sampling distribution. In particular we may use the policy π to restrict samples to situations that cause the neuron to spike to obtain P (s, νπ(s) = a1) = π(sa1) · P (νs). It follows that L(π) = Rπ = E(cid:104) (cid:98)L(π) = (cid:98)Rπ = 1 T1 R(s, a1, ν) (cid:12)(cid:12)(cid:12) π(sa1) · P (νs) (cid:105) (cid:88) R(cid:0)st, a1, νt {(st,νt)π(st)=a1} is the expected reward and (cid:1) is the empirical reward, where T1 is the number of spike produced by the neuron during [1, T ]. Let H ⊃ Hk = (cid:8)π : S → A s.t. π−1(a1) = k(cid:9) denote policies that spike for exactly k Theorem 3. situations. Given the setup above, with probability at least 1 − δ, (cid:115) (cid:12)(cid:12)(cid:12)Rπ − (cid:98)Rπ (cid:12)(cid:12)(cid:12) ≤ b · . By Stirling's approximation, log(cid:0)N (cid:19) (cid:18)N k Proof. Let N = S denote the number of possible situations. We put the uniform prior on Hk, so P (π) = 1 (N k ) − log P (π) = log (cid:1), and it follows that (cid:19) (cid:1) ≤ k log(cid:0) N·e ≤ k log (cid:18) + 1 log · k . N · e k = N · k N N k k S · eiu(π, a1) + 1 2T1 · eeiu(π,a1) + log 2 δ 2T1 The theorem follows since πu(a1) = k N and eiu(π, a1) = log N k . 12 (a) Effective information from spikes (b) Empirical reward after spiking Figure 1: Metabolically constrained Q-learning. As neurons learn, effective infor- mation and empirical reward increase in qualitatively the same way. Tighter metabolic constraints yield both higher effective information and greater rewards. 13 P(a1)=0.5P(a1)=0.3P(a1)=0.1P(a1)=0.5P(a1)=0.3P(a1)=0.1Rπ Figure 2: Concentration of reward. The x-axis lists situations, ranked by the empirical reward the neuron would receive if it spiked. Situations are grouped into two categories: the top 15%, which cause spikes, and the rest, which do not. The average and variance of the empirical reward in each category is displayed. Figure 3: Empirical versus expected reward. The normalized difference between ex- pected and empirical reward, plotted against effective information. 14 Empirical reward after spikingEmpirical reward after silence15 percentileμ = 0.21σ(cid:31) = .054μ = 0.89σ(cid:31) = .003EMPIRICAL REWARDSILENTSPIKE
1611.06197
1
1611
2016-11-18T18:53:45
An Empirical Study of Continuous Connectivity Degree Sequence Equivalents
[ "q-bio.NC", "cs.NE" ]
In the present work we demonstrate the use of a parcellation free connectivity model based on Poisson point processes. This model produces for each subject a continuous bivariate intensity function that represents for every possible pair of points the relative rate at which we observe tracts terminating at those points. We fit this model to explore degree sequence equivalents for spatial continuum graphs, and to investigate the local differences between estimated intensity functions for two different tractography methods. This is a companion paper to Moyer et al. (2016), where the model was originally defined.
q-bio.NC
q-bio
An Empirical Study of Continuous Connectivity Degree Sequence Equivalents Daniel Moyer, Boris A. Gutman, Joshua Faskowitz, Neda Jahanshad, and Paul M. Thompson Imaging Genetics Center, University of Southern California [email protected] Abstract. In the present work we demonstrate the use of a parcellation free connectivity model based on Poisson point processes. This model produces for each subject a continuous bivariate intensity function that represents for every possible pair of points the relative rate at which we observe tracts terminating at those points. We fit this model to explore degree sequence equivalents for spatial continuum graphs, and to inves- tigate the local differences between estimated intensity functions for two different tractography methods. This is a companion paper to [11], where the model was originally defined. Keywords: Human Connectome, Diffusion MRI, Non-Parametric Esti- mation 1 Introduction In the past decade, graph theoretic analyses have rapidly propagated through neuroimaging literature. Following advances in diffusion and functional MRI, the rise of connectomics has popularized the use of network representations of brain architecture and activity. Such analyses usually equate physical regions of the cortical surface with nodes in a graph, and use structural or function measurements as proxies for edge weights. Increased use of the network representation of brain connectivity has been accompanied by the use of network statistics in disease-oriented neuroscience; nodal measures such as modularity, centrality, and degree sequences are all com- mon descriptors for connectomes [2], and the effect of various diseases on these descriptors has been the focus of recent study [4]. The popularity of these mea- sures stems in part from their theoretical underpinnings. In particular, the distri- bution of nodal degrees has deep implications for the organization and topology of a random network. For brain networks the choice of a particular delineation of physical regions (parcellation) and thus the choice of nodes is non-trivial. Multiple studies have shown that the choice of parcellation influences the summary measures including nodal degree for both structural and functional networks [13,17,18]. It remains unclear which of the many parcellations is optimal, or whether or not a single overall "best" parcellation exists [12], assuming a criterion for quality can be agreed upon. Furthermore, the study of local or multi-scale grained phenomena often require specific resolutions and thus specific parcellations. This presents a dilemma to the community: do we choose the best parcellation for their specific study, or do we choose a parcellation that generalizes well to other literature? It is therefore valuable to explore alternative representations that avoid these issues. In particular, it is useful to construct representations of cortical connec- tivity that are independent of the choice of parcellation, yet still theoretically and computationally tractable both for estimation as well as statistical analy- sis, as well as retaining the ability to construct parcellation based connectivity. Furthermore, the exploration of continuous equivalents to currently popular net- work statistics and their similarities and/or deviations from current empirical observations is of interest. In the current work we explore the empirical properties of such a model, described in a companion paper [11]. We explore a continuous representation of cortical connectivity that describes the observation of white matter tract1 endpoints using a random process defined on the gray matter/white matter interface (the inner cortical surface). This form of connectivity generalizes tra- ditional connectomes to a parcellation-independent representation from which, given any particular parcellation, the discrete connectome may be recovered. We reproduce degree distribution results similar to those described by discrete representation analyses using our continuous model, and note the instances of discrepancy between the two. We further investigate differences between two tractography methods in the degree distribution (marginal intensity function). 2 Continuous Connectivity Model In order to ensure that the reader may understand the results in Section 3, we first review the theory and motivation behind this particular continuous connectivity model; more details can be found in [11]. In particular we introduce the framework itself and its general terminology, focusing on the key piece of our model, the Poisson point process and its intensity function. We then define the analogous statistic to degree for the continuous framework, which will serve as the empirical focus of the next section. 2.1 Model Description and Theoretical Discussion A point process is a random process in which collections of discrete points are generated randomly on a measurable space. The Poisson process is the most basic of these, assuming that these points are generated independently (i.e. the appearance of one point does not affect the probability of observing another) and with some relative rate proportional to an intensity function λ : Domain → R+. For any subset of the domain the distribution of the number of points in that 1 It is important to distinguish between white matter fibers (fascicles) and observed "tracts." Here, "tracts" denotes the 3d-curves recovered from Diffusion Weighted Imaging via tractography algorithms. subset is Poisson with parameter equal to λ integrated over the subset. This means the expected number of points is exactly that integral. We use this process to model tract endpoint locations on the cortical surface. Our domain is the connectivity space of the cortex (the product of the gray/white matter interface with itself) which is the set of all pairs of possible endpoints, and each point in this space corresponds with a pair of tract endpoints. In our context of connectomics, this framework produces two different struc- tures that are both analogous to connectivity. The first is the usual region-to- region connectivity, which is produced by measuring the expected tract count on a subset of the domain which is itself composed of all pairs of points in two sub- sets of the cortex. These cortical subsets are not necessarily disjoint. The second representation of connectivity is the intensity function λ. While the first is an aggregation of the second, it is important to separate them. The intensity func- tion has pointwise intensities which are not comparable with the second. The second representation generalizes the first over the choice of regions. Varying segmentation choices for the first representation can be compared in the context of the second. A more formal definition of the framework is as follows: Let Ω be union of two disjoint subspaces each diffeomorphic to the 2-sphere representing the cortical surface for a particular hemisphere, and assume that tracts randomly and independently intersect with this surface at exactly two points. Further consider the space Ω × Ω, which is the space of pairs of points on the cortical surface, in our case the space of all possible pairs of endpoints of tracts. We model connectivity as a function λ : Ω × Ω → R+ such that for any regions E1, E2 ⊂ Ω the number of observed tracts having one end point in E1 and the other in E2 is Poisson distributed with parameter (cid:90)(cid:90) C(E1, E2) = λ(x, y)dxdy. E1,E2 (1) This is exactly a Poisson point process of tract endpoints over Ω × Ω with λ is its intensity function. Note that for E1, E2 with non-trivial intersections this double counts the number of actual tracts observed in that intersection. If a tract has endpoint (x, y) ∈ E1 ∩ E2, then clearly there exists a tract (y, x) ∈ E1 ∩ E2. We define λ(x, y) for any particular (x, y) as the pointwise connectivity be- tween x and y. Though it is not technically required, we assume this pro- cess (characterized by λ) to be smooth everywhere and symmetric (λ(x, y) = λ(y, x)).. We further assume tracts are generated independently; more complex point process models relax this assumption, and a more general form of this model could be considered, in which λ is vector valued, and the Poisson dis- tribution may be substituted for any distribution parameterizable by integral terms. These models are unfortunately computationally intractable with current estimation methods, but are relevant to the proceeding discussion. We define a regional connectivity as the expected number of tracts between any two regions. In the Poisson case this is exactly C(E1, E2). For any parti- i Ei = Ω, this forms a connectivity matrix similar to tion (parcellation) P =(cid:83) the traditional connectomes (which provide connectivity for the P × P discrete space). From empirical data, a valid estimator is simply counting the number of tracts present between each pair of regions, as current methods dictate. Thus, continuous connectivity models of the form proposed here generalizes graphs generated by finite partitions of Ω (including overlapping paritions) for certain classes of edge weight distributions. An important summary statistic of traditional graphs is the degree sequence, defined as the collection of sums of each edge for each node (this counts each edge weight exactly twice, though for different terms in the sequence). A sim- ilar construct exists for the continuous connectome which we call the marginal connectivity, given by (cid:90) M (x) = λ(x, y)dy. Ω M (x) itself is defined over Ω. Using the assumption that λ is continuous, it can be shown fairly easily that M (x) is also continuous. Though not explored in this paper, another natural extension are marginal connectivities from a particu- E λ(x, y)dy, lar region. Choosing any region E, we can define them as ME(x) =(cid:82) which is the total connectivity from point x to the region E. 2.2 Estimation and Asymptotic Estimator Distribution A sufficient statistic for Poisson process models is the intensity function λ(x, y). Estimation of the function is non-trivial, and has been the subject of much study in the spatial statistics community [5]. We choose to use a non-parametric Kernel Density Estimation (KDE) approach due to an efficient closed (truncated harmonic) form for estimation. We first inflate each surface to a sphere and register them using a spherical registration (See section 3.1), assuming each hemisphere is disjoint with the other. Our domain is then Ω × Ω the product of spheres (S1 ∪ S2) × (S1 ∪ S2). The unit normalized spherical heat kernel is a natural choice of kernel for S2. We use its truncated spherical harmonic representation [3], defined as follows for any two unit vectors p and q on the 2-sphere: H(cid:88) h Kσ(p, q) = 2h + 1 4π exp{−h(h + 1)σ}P 0 h (p · q) h is the hth degree associated Legendre polynomial of order 0. Note that Here, P 0 the non-zero order polynomials have coefficient zero due to the radial symmetry of the spherical heat kernel [3]. However, since we are estimating a function on Ω × Ω, we use the product of two heat kernels as our KDE kernel κ. For any two points p and q, the kernel value associated to a end point pair (x, y) is κ((p, q)(x, y)) = Kσ(x, p)Kσ(y, q). tion. This is exactly λ(p, q) = κ((p, q)D) =(cid:80) We then apply this kernel to our data in order to recover the intensity func- (xi,yi)∈D Kσ(xi, p)Kσ(yi, q) where λ denotes our estimation of λ, the true intensity. Since storing the exact func- tional form is expensive, in practice we evaluate this at each vertex of the cortical surface triangulated mesh. It should be noted that while the end result of this process is an array of values corresponding to a triangulated mesh of the cortical surface (which can be inflated into a sphere), this is not equivalent to a high resolution discrete con- nectome except in the roughest terms. Each point here represents a pointwise evaluation of the density function, and the difference between adjacent mesh vertices is bounded (since the domain is compact, we can show that the true intensity function must be absolutely continuous if it is continuous at all). The discrete connectome, no matter how small the region, represents an areal con- nectivity value, and in most cases caries no geometric information, much less a smoothness condition. The product kernel complicates the asymptotic analysis of this estimator. Unfortunately there is no true closed form for the heat kernel on the sphere [10], and approximations include terms that are not consistent with traditional kernel bandwidth analysis (which usually is univariate, and unbounded). On the other hand, it can be shown that the heat kernel using these approximate forms will be consistent with extant first and second moments [8]. 3 Empirical Results 3.1 Pre-processing Our data are comprised of 731 subjects from the Human Connectome Project S900 release [16]. We used the minimally preprocessed T1-weighted (T1w) and diffusion weighted (DWI) images rigidly aligned to MNI 152 space. Briefly, the preprocessing of these images included gradient nonlinearity correction (T1w, DWI), motion correction (DWI), eddy current correction (DWI), and linear alignment (T1w, DWI). We used the HCP Pipeline (version 3.13.1) FreeSurfer protocol to run an optimized version of the recon-all pipeline that computes surface meshes in a higher resolution (0.7mm isotropic) space. Tractography was conducted using the DWI in 1.25mm isotropic MNI 152 space. Probabilistic streamline tractography was performed using Dipy's Local- Tracking module [7]. To model the fiber distribution at each voxel, Dipy's im- plementation of constrained spherical deconvolution (CSD) [15] was used with a spherical harmonics order of 8. Tractography streamlines were seeded at 2 ran- dom locations in each voxel labeled as likely white matter via the segmentation maps generated by FSL's FAST. Streamline tracking followed directions ran- domly in proportion to the orientation function at each sample point at 0.5mm steps, starting bidirectionaly from each seed point. Streamlines were only re- tained if longer than 5mm and both ends terminated in voxels likely correspond- ing to grey matter according to Dipy's implementation of ACT[14]. An additional tractography was computed in the same manner, but replacing the CSD fit for a diffusion tensor (DTI) fit. Fig. 1. Here we see the estimated density of sampled marginal connectivity functions. The top figure is the direct plot, and the bottom figure is the log-log plot of the same. We then fit 5 separate scales of continuous connectivity models, varying σ as to capture differing scales of spatial patterns. We use each σ = {0.01,0.005, 0.001, 0.0005, 0.0001}, for both DTI and CSD . We then integrate numerically each resulting intensity function for each subject, forming two marginal connec- tivity functions for each subject. A paired t-test was then performed at each of the 20484 vertex across subjects (the surface is subsampled for intensity function computation). Multiple comparison correction was applied before inference us- ing the Bonferroni correction for 20484 hypotheses. The continuous assumption makes this correction overly conservative; the test points are no longer inde- Fig. 2. In this plot we see p-values for the CSD-DTI test. Blue denotes CSD > DTI by as significant margin, while Red denotes DTI > CSD by a significant margin. Gray areas denote regions that were unable to reject the null hypothesis of "no difference". From top to bottom we have σ = {0.01,0.005, 0.001, 0.0005, 0.0001}, and from left to right we have the left hemisphere and right hemisphere. pendent due to the smoothness of the signal. Test statistics were evaluated at α = 0.05. We also estimate continuous one dimensional densities for samples from the average marginal connectivity function's distribution of values. 3.2 Degree Sequences Much sensation in the past 20 years has focused on complex networks and the distribution of their degree sequences [1]. We here display estimates for a similar measure for sampled marginal connectivity functions. Multiple studies have pro- vided evidence of small-worldness or scale-free properties in connectomes, both functional and structural [6,2], sometimes making conflicting claims. While we here make no rigorous claim about the fit of one model or another, we do note the dissimilarity to strictly power law distributions, and their similarity to ge- ometrically constrained communication networks, particularly ad hoc computer networks [9]. (This is qualitative comparison only, and not backed by quanti- tative observation of ad hoc networks). The plots in Fig 1 show clear second modes, and their log plots are slightly non-linear. 3.3 DTI-CSD Difference As shown in Figure 2, there are clear differences recovered by the continuous connectivity model between the CSD and DTI tractography algorithms. This not only illustrates use of a continuous model, but is important for understanding the different trade offs between tractography algorithms. In general DTI is much less flexible than probabilistic CSD, but also fit orders of magnitude faster. Since each connectome is normalized, significant areas are proportionally more explored by a particular algorithm. The authors would also like to reiterate the value and importance of multiple test correction in large hypothesis set situations such as this, due to the large number of sample points tested. DTI appears to concentrate on motor and somatosensory cortices and their corresponding tracts, while the CSD model appears to find more tracts in the temporal lobe. This is not to say that CSD is missing the motor tracts or that DTI does not have any temporal lobe tracts, but that the relative concentrations are shifted. It is also interesting to note that several small regions break this trend, particularly for DTI concentrations in the inferior temporal and temporal- frotal regions for σ ≤ 0.001. While this work is preliminary, this demonstrates some of the advantages of continuous connectivity models; while region specific models may not be able to resolve such small differences, particularly those surrounded by significant regions of the other density, the continuum model allows very local differences to be detected, as well as large, non-convex regions (such as those in the parietal/frontal regions for DTI, with σ = 0.01). This also illustrates one unfortunate issue with the continuum model: Visual- ization of the full connectivity is difficult (a four dimensional continuum usually embedded in six dimensions and that is not a 4-sphere). In future work we plan on addressing this issue and developing methods to visualize portions at a time. However, we also believe this speaks to the possible issues when visualizing the discrete networks. 4 Conclusion In the present work we have described a parcellation free connectivity model. We further used this model to explore degree sequence equivalents for spatial continuum graphs, and to investigate pointwise differences in these functions for two different tractography methods. We believe that parcellation based networks are critical to the exploration of the cortical landscape, but that the development of more general methods such as the one presented here is also vital to expanding the set of testable questions in neuroscience, and improving the answers provided by neuroimaging. References 1. Barab´asi, A.L.: Scale-free networks: a decade and beyond. science 325(5939), 412 -- 413 (2009) 2. Bullmore, E., Sporns, O.: Complex brain networks: graph theoretical analysis of structural and functional systems. Nature Reviews Neuroscience 10(3), 186 -- 198 (2009) 3. Chung, M.K.: Heat kernel smoothing on unit sphere. In: Biomedical Imaging: Nano to Macro, 2006. 3rd IEEE International Symposium on. pp. 992 -- 995. IEEE (2006) 4. Crossley, N.A., Mechelli, A., Scott, J., Carletti, F., Fox, P.T., McGuire, P., Bull- more, E.T.: The hubs of the human connectome are generally implicated in the anatomy of brain disorders. Brain 137(8), 2382 -- 2395 (2014) 5. Diggle, P.: A kernel method for smoothing point process data. Applied Statistics pp. 138 -- 147 (1985) 6. Eguiluz, V.M., Chialvo, D.R., Cecchi, G.A., Baliki, M., Apkarian, A.V.: Scale-free brain functional networks. Physical review letters 94(1), 018102 (2005) 7. Garyfallidis, E., et al.: Dipy, a library for the analysis of diffusion MRI data. Front. Neuroinform 8(8) (2014) 8. Hall, P., Watson, G., Cabrera, J.: Kernel density estimation with spherical data. Biometrika 74(4), 751 -- 762 (1987) 9. Hekmat, R., Van Mieghem, P.: Connectivity in wireless ad-hoc networks with a log-normal radio model. Mobile networks and applications 11(3), 351 -- 360 (2006) 10. Lafferty, J., Lebanon, G.: Diffusion kernels on statistical manifolds. Journal of Machine Learning Research 6(Jan), 129 -- 163 (2005) 11. Moyer, D., Gutman, B., Faskowitz, J., Thompson, P.M.: A continuous model of cortical connectivity. In: MICCAI 2016. Springer (2016) 12. de Reus, M.A., Van den Heuvel, M.P.: The parcellation-based connectome: limita- tions and extensions. NeuroImage 80, 397 -- 404 (2013) 13. Satterthwaite, T.D., Davatzikos, C.: Towards an individualized delineation of func- tional neuroanatomy. Neuron 87(3), 471 -- 473 (2015) 14. Smith, R.E., et al.: Anatomically-constrained tractography: improved diffusion MRI streamlines tractography through effective use of anatomical information. NeuroImage 62(3), 1924 -- 1938 (2012) 15. Tournier, J.D., Yeh, C.H., Calamante, F., Cho, K.H., Connelly, A., Lin, C.P.: Re- solving crossing fibres using constrained spherical deconvolution: validation using diffusion-weighted imaging phantom data. NeuroImage 42(2), 617 -- 625 (2008) 16. Van Essen, D.C., Smith, S.M., Barch, D.M., Behrens, T.E., Yacoub, E., Ugurbil, K., Consortium, W.M.H., et al.: The wu-minn human connectome project: an overview. Neuroimage 80, 62 -- 79 (2013) 17. Wang, J., et al.: Parcellation-dependent small-world brain functional networks: A resting-state fMRI study. Human Brain Mapping 30(5), 1511 -- 1523 (2009) 18. Zalesky, A., et al.: Whole-brain anatomical networks: does the choice of nodes matter? NeuroImage 50(3), 970 -- 983 (2010)
0810.0029
3
0810
2010-03-19T23:12:26
Topological Effects of Synaptic Time Dependent Plasticity
[ "q-bio.NC", "cond-mat.dis-nn", "math-ph", "math-ph", "q-bio.TO" ]
We show that the local Spike Timing-Dependent Plasticity (STDP) rule has the effect of regulating the trans-synaptic weights of loops of any length within a simulated network of neurons. We show that depending on STDP's polarity, functional loops are formed or eliminated in networks driven to normal spiking conditions by random, partially correlated inputs, where functional loops comprise weights that exceed a non-zero threshold. We further prove that STDP is a form of loop-regulating plasticity for the case of a linear network comprising random weights drawn from certain distributions. Thus a notable local synaptic learning rule makes a specific prediction about synapses in the brain in which standard STDP is present: that under normal spiking conditions, they should participate in predominantly feed-forward connections at all scales. Our model implies that any deviations from this prediction would require a substantial modification to the hypothesized role for standard STDP. Given its widespread occurrence in the brain, we predict that STDP could also regulate long range synaptic loops among individual neurons across all brain scales, up to, and including, the scale of global brain network topology.
q-bio.NC
q-bio
A Theory of Loop Formation and Elimination By Spike Timing-Dependent Plasticity James Kozloski and Guillermo A. Cecchi Computational Biology Center, IBM Research Division, IBM T.J. Watson Research Center, Yorktown Heights, NY 10598 (Dated: October 25, 2018) Abstract (FULL PAPER IN FRONTIERS IN NEURAL CIRCUITS, 2010) We show that the local Spike Timing-Dependent Plasticity (STDP) rule has the effect of regulating the trans-synaptic weights of loops of any length within a simulated network of neurons. We show that depending on STDP's polarity, functional loops are formed or eliminated in networks driven to normal spiking conditions by random, partially correlated inputs, where functional loops comprise weights that exceed a non- zero threshold. We further prove that STDP is a form of loop-regulating plasticity for the case of a linear network comprising random weights drawn from certain distributions. Thus a notable local synaptic learning rule makes a specific prediction about synapses in the brain in which standard STDP is present: that under normal spiking conditions, they should participate in predominantly feed-forward connections at all scales. Our model implies that any deviations from this prediction would require a substantial modification to the hypothesized role for standard STDP. Given its widespread occurrence in the brain, we predict that STDP could also regulate long range synaptic loops among individual neurons across all brain scales, up to, and including, the scale of global brain network topology. 0 1 0 2 r a M 9 1 ] . C N o i b - q [ 3 v 9 2 0 0 . 0 1 8 0 : v i X r a 1 I. INTRODUCTION Connections between individual neurons in the brain are constrained first by the spa- tial distribution of axons and dendrites within the neuropil [1][2]. Global brain networks comprise dense connections within tissues, the gross structures in which these tissues are embedded, and the bidirectional long-range projections joining these structures. The topol- ogy of these networks is not yet fully specified at the level of microcircuitry, however [24]. One theoretical constraint on this level of organization, the "no strong loops hypothesis," considered only developmentally determined area to area connectivity patterns to imple- ment its specific neuron to neuron network topological constraint [3]. While local synaptic modifications are known to directly shape the pattern of connectivity in local neural tissue and thus local microcircuit topology [4], our understanding of global brain network topology still derives largely from this developmentally patterned, area to area connectivity. Further- more, measuring simultaneously the relative strengths of specific microcircuit connections remains technically challenging, and virtually impossible for even medium sized (100 − 200 neurons, 0.05 − 0.1 mm) microcircuits. For these reasons, it is not yet known how large scale, long range microcircuit topology and the computation it supports emerges through synaptic modifications in the brain. We wondered whether a synaptic modification commonly observed in local circuit prepa- rations and widely hypothesized to shape local dynamics in brain structures, STDP [5], could be analyzed to yield an understanding of what topology it predicts for microcircuits of any scale. The STDP model is a departure from traditional Hebbian models of learning, which state that neurons that fire action potentials together will have their interconnections strengthened. Instead, STDP takes into account the particular temporal order of pre- and post-synaptic neuronal firing [6], such that the rule modifies synapses anti-symmetrically, depending on whether the pre- or post-synaptic neuron fires first (Fig. 1). The basic ques- tion we then aimed to answer is: what is the influence of this anti-symmetry on brain microcircuit topology? Consider first if a pre-synaptic "trigger" neuron causes a post-synaptic, first-order "fol- lower" neuron to fire. If this follower makes a direct feedback connection onto the trigger, the feedback connection will be weakened, since the spike generated by the follower will arrive at the follower-trigger synapse immediately after the trigger neuron's backward propagating 2 action potential (Fig. 1). The principle that STDP is suitable for eliminating strong recur- rent connections between two neurons was originally proposed by Abbott and Nelson [7]. Here we expand on the principle with the observation that it holds for all polysynaptic loops connecting triggers and followers: if some nth-order follower's action potential produces in the original trigger a subthreshold potential after the trigger has fired, the functional loop will be broken by spike-timing dependent synaptic weakening of the feedback connection. With this intuition, we set out to prove analytically and by means of numerical simulation that network topology, and specifically the occurrence of functional loops in highly connected networks, is directly and necessarily regulated by STDP. This theory paper provides clear predictions about STDP's effect on neural circuit topol- ogy. The proof and simulations dictate strong constraints on local and long range micro- circuit connectivity. We propose that if these constraints are not obeyed by real neural circuits, the hypothesis that standard STDP shapes the structure and function of real ner- vous systems must be revised. Our approach suggests that similar analyses of other learning rules may impose similar constraints on neural circuit topology and that the hypothetical significance of these rules may similarly be tested. II. RESULTS A. A proof of STDP as a form of loop-regulating plasticity First, we represent STDP acting on a weight w associated with the connection between two neurons and their output variables x(t) to y(t), in the adiabatic approximation (i.e. small learning rate), as: (cid:90) ∞ −∞ where Cxy(t) =(cid:82) x(t(cid:48)−t)y(t(cid:48)) is the correlator, and S(t) is the anti-symmetric STDP update ∆wxy ∝ Cxy(t)S(t)dt function, S(t < 0) = exp(λt), S(t > 0) = − exp(−λt). Consider this function operating over connections within a linear network driven by uncorrelated Gaussian inputs, ξ, such that (1) x(t) = W x(t) + ξ(t), where x is a vector of activities with components denoted by xi, the weight connection matrix has components Wij, and the input satisfies (cid:104)ξi(t)ξT j (t + τ )(cid:105)t = σ2δ(τ )δij. We show (see Appendix A 1) that the learning rule defined in Eq. 1 results in an update for the network weight matrix of the form ∆W = ∆W (W, τ, C0) where τ is the 3 time constant of the STDP's exponential, and C0 is the instantaneous correlator C(0). This update rule influences global network topology in a very specific way. To formalize our original intuition analytically, consider a linear network with only ex- citatory connections, such that the dynamics may be expressed as x = W x = (−I + A)x, where Aij ≥ 0 is the network connectivity matrix (comprising the off-diagonal elements of the weight matrix and zeros on the diagonal), and −I represents a self-decay term. Next, we introduce a "loopiness" measure that estimates the strength of all loops of all sizes that k tr (cid:2)Ak(cid:3). The function tr [·] stands for the trace operation; occur in the network, El =(cid:80)∞ 1 k=1 this operation, when acting on the k-exponentiation of the adjacency matrix (comprising ones and zeros, where the nonzero entry aij represents a connection from network node i to network node j), counts the total number of closed paths of length exactly equal to k, i.e. k-loops [25]. When applied to the network connectivity matrix A, the operation counts loops weighted by the product of the synaptic strengths of the looping connections, resulting in a slightly different, but still useful, measure of loopiness. It is possible, however, to reduce this measure without actually regulating topology by simply reducing the weights of all connections. A topological therefore include a penalty to the weights' vanishing; we choose − 1 the transpose operation) which, for weighted graphs, measures the sum of the squares of all loopiness measure should 2tr (cid:2)AAT(cid:3), (T stands for network weights (and for a binary graph, counts the number of links). We then define the total topological loopiness as: E = ∞(cid:88) k=1 1 k tr (cid:2)(A)k(cid:3) − 1 2 tr (cid:2)AAT(cid:3) (2) We showed computationally that for every weight matrix examined, each drawn from certain random distributions (see Appendix A 1), the change in this energy as a function of the evolution of the network under STDP, ∆E ∼ tr (cid:2)∂AE∆AT(cid:3), is strictly semi-negative, and therefore STDP necessarily regulates this measure, resulting in a decrease in topological loopiness. We therefore use the term STDP and "loop-regulating plasticity" interchangeably throughout. 4 B. Loop-regulating plasticity in a network of simulated neurons What are the effects of this form of plasticity on network topology (and specifically on the number of functional loops) in nonlinear networks, such as those found in neural microcircuits? Because our proof of STDP as a form of loop-regulating plasticity applies only to linear networks or nonlinear networks that may be linearized, we aimed to show, using simulation, that the same principle extends to a biologically relevant, nonlinear regime. We replicated the simulation of Song and Abbott [8], extending it in three ways (see IV A). First, we created a network of 100 neurons, each receiving excitatory synapses from all other 99 "intra-network" input sources and from 401 randomly spiking "extra-network" input sources selected at random from 2,500 homogeneous Poisson processes. All excitatory synapses underwent STDP. Second, we provided 250 inhibitory synapses to each neuron, sampled from 1,250 spiking sources; the inhibitory inputs modeled fast local inhibition to the network using inhomogeneous Poisson processes with rates modulated by the instantaneous aggregate firing rate of the network. Third, we explored four different forms of STDP update [9] and observed robust loop-regulating plasticity for each; the results presented here used the STDP update rule of Gutig et al. [10]. We initialized our network with maximum extra-network weights, and intra-network weights at half maximum. This caused the network to spike vigorously when extra-network inputs became active, but spike rates were limited by the fast local-inhibition. After 20 sec- onds of simulated network activity, STDP had a profound effect on topological loopiness as defined in Eq. 2, measured over loops of length 2 ≤ k ≤ 100 for convenience (Fig. 2A). We counted the number of closed, functional loops of varying length using tr (cid:2)(cid:100)A(cid:101)k(cid:3), where (cid:100)A(cid:101) was constructed by applying a sliding threshold to the network connectivity matrix (Fig. 2B). We compared this quantity to the same, measured for a randomized network, con- structed by randomly reassigning weights from the learned weight distribution to synapses in the network (see IV B). These results are representative of all loop lengths measured (2 ≤ n ≤ 100) and show that as the weight threshold grows, the number of closed, func- tional loops in the STDP-learned network decreases more than in the randomized network. This form of loop-regulating plasticity can therefore be described as loop-eliminating. 5 C. The effect of synaptic delays on loop-regulating plasticity We wondered what effect synaptic delays would have on this result, since we expected follower feedback spikes to cause less anti-loop learning as they fell further from the zero time difference maxima in the STDP update function. We also wondered if the decrease in the number of functional loops compared to a randomized network also applied to unique loops, in which no neuron is traversed more than once. We therefore sampled the number of unique, functional loops through networks simulated with synaptic delays from 0.1 to 4.0 milliseconds. We constructed one million random paths of length k − 1 for each loop length 2 ≤ k ≤ 25, and for the learned and randomized networks (see IV B). We searched for each path across all networks studied, and if the path and the kth link completing the functional loop existed in the network, we counted it for that network (see IV B). The result is similar to that for closed loops, and, as expected, longer synaptic delays resulted in an exponential decrease in the number of loops as a function of loop length that deviated less from the same function for randomized networks, indicating weaker loop-regulating plasticity (Fig. 2C). D. Network in-hubs, out-hubs, and loop-regulating plasticity Next, we asked if other topological measures of the STDP-learned networks may be cor- related with our observation of STDP's effect on loopiness, since many different topological properties might coincide with or support this effect. For example, one means to create networks poor in loops is to ensure that nodes in the network are either "out-hubs" or "in- hubs," but not both [11]. An out-hub in a network of neurons has many strong postsynaptic connections but few strong presynaptic connections, and an in-hub has many strong presy- naptic connections but few strong postsynaptic connections. We applied a sliding threshold to the network connectivity matrix learned by STDP, and examined the manifold, colored according to each applied threshold, which correlated in-degree versus out-degree for each neuron in our network. This showed a clear inverse relationship between in- and out-degrees that varied in form with weight threshold (Fig. 3A). In contrast, by examining the in- degree from extra-network inputs, we found a positive correlation (Fig. 3B), indicating that out-hubs were more likely to be in-hubs within the larger extra-network topology, and that in-hubs in our network were more likely to receive only the weakest extra-network inputs. 6 E. Reverse STDP restored loops after loop-eliminating plasticity Beyond these standard topological analyses, we also examined biological properties of the network. We measured total synaptic input as a function of total synaptic output for all neurons in the STDP-learned network. In the same experiment, we asked if reversing the polarity of the standard STDP function might undo the effects of loop-regulating plastic- ity that results from standard STDP, since under this "reverse" condition, follower spikes would cause strengthening of closed-loop feedback connections. This reversal of polarity is biologically relevant, since it occurs at the synaptic interface between major brain struc- tures such as neocortex and striatum [12], arises specifically at synapses between certain cell types, and is controlled by cholinergic and adrenergic neuromodulation, for example in the neocortical microcircuit [13]. We found the same inverse relationship between in-degree versus out-degree for each neuron in our network (Fig. 4A, green markers), as well as an in- verse relationship between total synaptic input and output following 1.5 seconds of standard STDP (Fig. 4B, green markers). These effects contributed to a reduction in the number of closed loops (Fig. 4C, depicted as in Fig. 2B), and each of these relationships could be largely abolished by 3 to 5 additional seconds of reverse STDP (Fig. 4, red markers), in contrast to 3 to 5 additional seconds of standard STDP (Fig. 4, blue markers), which strengthened them. We also found the same positive correlation between in-degree from extra-network inputs and out-degree within the network (Fig. 4D) and between total extra- network synaptic input and total intra-network synaptic output (Fig. 4E). This effect was also largely abolished by 3 to 5 seconds of reverse loop-regulating plasticity, but reinforced by 3 to 5 seconds of standard loop-regulating plasticity. F. Dynamical effects of loop-regulating plasticity What are the consequences of this form of network plasticity beyond topology? In the case of a linear network, reducing the number of loops implies more stable dynamics. Consider the stability of the unforced system (cid:126)x(t) − W (cid:126)x(t) = 0; the eigenvalues λ of W = −I + A can be expressed as: (3) ∞(cid:88) tr (cid:2)Ak(cid:3) 1 k − N(cid:88) log λi = i=1 k=1 7 [14], which emphasizes the contribution of loops to system instability. Such a simple ob- servation, however, does not make clear predictions about the effects of loop-regulating plasticity on nonlinear neural circuit function. We were surprised to find that raster plots of network spiking activity, when sorted according to certain topological metrics (e.g., the sum of extra-network input weights, the sum of intra-network output weights, in-degrees, out-degree) consistently revealed network events that originate with weak synchronization among out-hubs, followed by strong synchronization among in-hubs (Fig. 5A, top), across 8 independent simulations of the phenomenon. This effect was altered by randomizing the intra-network weights, such that synchronization events became stronger, more frequently global, and more frequent among out-hubs alone (Fig. 5A, bottom). Peri-event time his- tograms constructed across 8 independent simulations reveal this same effect (Fig. 5B, left panels), with synchronization arising strongly among in-hubs after weak out-hub activation in the STDP-learned network, and globally in the randomized network (see IV B for a de- scription of how network events were detected). In the STDP-learned network, both in-hubs and out-hubs sustain spike rates ranging from 4−9 Hz that are not correlated with in-degree, whereas in the randomized network, spike rates range more broadly (3 − 16 Hz) and are highly correlated with in-degree (Fig. 5B, right panels). We examined the summed network peri-event time histograms for the STDP-learned network and for networks that underwent randomization of their intra-network weights, their extra-network weights, or both (Fig. 5C, top), across 8 simulations for each condition. The resulting distributions, as well as a pooled distribution of times from all randomized networks each differed from each other based on paired Kolmogorov-Smirnov tests (P ≈ 0). To examine what properties of these distributions distinguished them, we measured kurtosis and skew for each distribution from each simulation, and compared the distributions of kurtosis and skew measures between each group. Kurtosis differed significantly between the STDP-learned network topology and the randomized topologies (intra-network randomized, extra-network randomized, both randomized, Fig. 5C, bottom). Results from unpaired t-tests of the distribution of skew and kurtosis measurements between 4 simulation conditions (calculated separately for each of the 8 simulations) are shown in table inset (Fig. 5C, bottom). These effects indicate that the effect of standard loop-regulating plasticity is to generate network topologies that support network events with greater spread and sharper peaks in time. 8 III. DISCUSSION Based on our simulations and analytical results, we propose that standard STDP must produce a network topology in real neural tissues that is conspicuously poor in both closed and unique loops, and that it will segregate neurons into out- and in-hubs to achieve this. Such a prediction can easily be tested by analyzing correlations between the number of functional input connections and the number of functional output connections made by neurons recorded during multi-patch clamp experiments in a structure in which STDP has been observed (e.g., [4], [15]). Our theory predicts this correlation should be negative. The network that emerges in such tissues will organize its relationship to inputs from other structures in an orderly fashion, making local out-hubs the primary target for long range inputs, and thus establishing a feed forward relationship between the network and its pool of inputs. In a larger system, we anticipate that local in-hubs would become long range outputs. This prediction may also be tested by correlating the local topological relationships of a neuron with its identified role as either an input, output, or interneuron within that structure. We also make a clear prediction for the effect of synaptic delays on modulating the topological effects of STDP. Correlations between these delays and functional connectivity data from multi-patch clamp recordings from connected neurons undergoing STDP can also be measured to determine if synaptic delays predict the strength of reciprocal connections. Furthermore, we observe that, at interfaces between brain structures where STDP is reversed by neuromodulation, circuit dynamics can be predicted from the expected change in network topology. For example, changes in STDP at the cortico-striatal synapse [16] resulting in reverse STDP would favor the emergence of strong cortico-striatal-thalamocortical loops resulting in oscillations in this circuit. Interestingly, the depletion of loops and the separation of nodes into out-hubs and in-hubs has been recently reported in a variety of complex biological systems, including functional networks at the level of spatio-temporal resolution of fMRI, and the neural network of C. Elegans [11], suggesting a general principle of organization and dynamical stability for entire classes of functional networks. These observations suggest quantitative measurements of topology in vertebrate microcircuits could produce similarly interesting results. It has been observed in local circuit preparations that a bias exists among layer 5 pyra- 9 midal neurons of rat neocortex towards strong reciprocal connectivity [4] [15], and towards looping motifs among triplets of this neuronal class [15]. Furthermore, cyclic connections are strongest among those neurons connected by the strongest synaptic weights. These same neurons also exhibits STDP at the excitatory synapses that join them [5]. Given our analy- sis, it is now clear that these observations contradict each other; specifically, we have shown that standard STDP under normal spiking conditions with random uncorrelated inputs is loop-eliminating. Therefore, other mechanisms and constraints than those we have analyzed must be at play. Consider the case of networks that have recently spiked at abnormal rates, either due to increased excitability within the network (e.g., due to injury, epilepsy, etc.) or due to otherwise elevated extra-network inputs. If the majority of post-synaptic potentials imme- diately cause action potentials, standard STDP may have the effect of strengthening loops. Also, a network driven by highly and specifically correlated inputs may spike in temporal patterns conducive to loop-strengthening by STDP (a hypothesis we are currently studying). Finally, as we have shown, a spiking network that has recently experienced a reversal of the polarity of STDP will also show an increase in the number of loops observed. Clearly more experiments and observation would be required in order to confirm or rule out each of these mechanisms. We observe that network activity propagates smoothly through the feed-forward topology generated by STDP (Fig. 5A, top panel) without segregating neurons by average spike-rates (Fig. 5B, upper right panel). The effect on global brain function of such properties would include stable average firing rates shared among all neurons, regardless of their topological position, and robust signal propagation, similar to "synfire chains" [17, 18]. Finally, our theory holds that the reversal of STDP's polarity represents a local switch for the modi- fication of both global brain network topology and global brain dynamics. Thus sources of modulation [19] that accomplish this reversal locally are in fact regulating global brain function by means of this switch. 10 IV. METHODS A. Simulation The simulation methods of Song and Abbot [8] were used to simulate each neuron in our 100 neuron network. We observed the reported topological results in each simulation after 10 seconds of network activity. In some cases, we performed additional, longer simulations of network activity to explore the convergence and stability of these topological measures under various conditions. Briefly, each neuron model was integrate-and-fire, with membrane dt = Vrest − V + gexc(t)(Eexc − V ) + ginh(t)(Einh − V ), potential determined as in [8], by τm with τm = 20ms, Vrest = −60mV, Eexc = 0mV, Einh = −70mV, Vthresh = −54mV, and Vreset = −60mV. dV The synaptic conductances gexc and ginh were modified by the arrival of a presynaptic spike, as in [8], such that gexc(t) → gexc(t) + ¯ga, and ginh(t) → gexc(t) + ¯ginh. dt = −ginh, with absence of a spike, these quantities decay by τexc τexc = τin = 5.0, ¯gin = 0.015, 0 ≤ ¯ga ≤ ¯gmax, and ¯gmax = 0.01. We initialized all elements ¯ga to different values for intra-network (¯ga = 0.005) and extra-network (¯ga = 0.01) inputs. dt = −gexc, and τinh In the dgexc dginh For extra-network inputs, excitatory homogeneous Poisson spike trains were generated at a rate of 20 Hz. Inhibitory, inhomogeneous Poisson spike trains that model fast local inhibition were generated at a rate rmin ≤ rinh ≤ rmax, where rmin = 5 Hz, and rmax = 1000 Hz. On each time step, dt = 0.1 ms, rinh is incremented by an amount proportional to the fraction, γ, of network neurons that spiked during that timestep, and decays with a time drinh dt = − [rinh + (rmax − rmin)γ]. After this update, if rinh exceeds constant, τr, such that τr rmax, rinh → rmax. For all simulations we report here, the STDP update rule for a synapse from neuron j to neuron i was ¯ga(i, j) = ¯ga(i, j) + ¯ga(i, j)µM (i) for synaptic potentiation, and ¯ga(i, j) = ¯ga(i, j) + (¯gmax − ¯ga(i, j))µP a(i, j), for synaptic depression, µ = 0.1 (¯ga is maintained in the interval [¯gmin, ¯gmax]). As in [8], M (i) and P (i, j) decay exponentially, such that τ− dM dt = −M (i) and τ+ dt = −Pa, τ+ = τ− = 20. Also as in [8], M (i) is decremented by A− every time a neuron i generates an action potential, A− = 0.00035, and Pa(i, j) is incremented dPa by A+ every time a synapse onto neuron i from neuron j receives an action potential, A+ = 0.00035. This update rule effectively implements the asymmetric function of STDP 11 (see Fig. 1). B. Analysis To randomize our networks for analysis (Figs. 2B, 2C, 4C, 5), we created a random sequence of indices ranging uniquely from 1 to n, where n was the number of off diagonal elements in our network's weight matrix. We used these indices to shuffle uniquely the positions of all off diagonal elements in the matrix, thus preserving the network's learned weight matrix, while destroying its learned topology. We chose to sample unique loops (Fig. 2C) in the networks rather than enumerating them, since for long loops (k > 20) the number of possible paths to search would exceed 1020. We therefore constructed one million random paths of length k − 1 for each loop length. We term these paths "unique" because we sampled units within each without substitution (i.e., no sub-loops were sampled in which a single node is traversed more than once). However, we allowed that each path could be represented more than once in the one million constructed paths. In fact, for the shortest paths constructed (k ≤ 3) this was necessarily the case since the total number of possible unique paths is less than one million. We sampled unique loops from adjacency matrices constructed such that the weight threshold produced a matrix that was precisely half-full (for 100 neurons, a matrix with 5, 000 ones and 5, 000 zeros). For 4.0, 2.0, 1.0, 0.5, and 0.1 millisecond delays, these thresholds were 0.0032, 0.0030, 0.0033, 0.0037, and 0.0046. In this way, we controlled across experiments for varying weight distributions, and sampled the same number of links for loops across all experiments. Counts were compared against the randomized network constructed from 5, 000 links. To detect network events (Fig. 5B, 5C), we employed the method described by Thivierge and Cisek [20]. Briefly, for each simulation in which network events were detected, we generated spike trains at 1 msec resolution for each neuron, equal in duration to the time series analyzed and comprising the same number of spikes as observed for that neuron. We constructed network spike time histograms of these spike trains across all neurons using a bin width of 10 msec. We then determined a threshold for the network equal to a count which 5% of these bins exceeded. Thresholds were determined 1000 times for each network and each simulation, and the mean of these 1000 values used as the threshold above which 12 network events were detected in network spike time histograms for each simulation. 13 Appendix A: Supplementary Information 1. Update rule for the weight matrix The classical definition of Hebbian learning for the weight w connecting two dynamical variables x(t) and y(t) can be written, in its simplest form, as: ∆W = ηCxy (cid:90) ∞ Cxy = −∞ x(t)y(t)dt (A1) (A2) Where η is the learning constant, which for exposition's sake will be set to 1. It is important, however, to keep in mind that in order to write Eqs. A1-A2 we are assuming an adiabatic approximation, i.e. the learning is small enough that the system can be considered to be in steady-state for the purpose of computing the correlation. A natural extension of Eqs. A1-A2 is to introduce time, i.e. to consider delayed as well as instantaneous correlations: ∆W ∼ Cxy(t)S(t)dt (cid:90) ∞ (cid:90) ∞ −∞ Cxy(t) = −∞ (cid:48) − t)y(t (cid:48) x(t (cid:48) )dt (A3) (A4) (A5) (A6) It is assumed that the time-dependent weight function vanishes for long delays, limt→±∞ S(t) = 0; the classical learning rule is recovered when S(t) = δ(t). If, as ex- perimental results describing STDP strongly suggest [5], the weight function displays strict temporal anti-symmetry, i.e. S(t) = −S(−t), then (cid:90) ∞ (cid:90) 0 ∆W ∼ (cid:90) ∞ −∞ ∆W ∼ Cxy(t)S(t)dt + Cxy(t)S(t)dt 0 [Cxy(t) − Cxy(−t)] S(t)dt A multi-dimensional linear system driven by uncorrelated input can be described as: 0 x(t) = W x(t) + ξ(t) (A7) where each unit is independently subject to Gaussian white noise ξ(t), a vector whose components satisfy (cid:104)ξi(t)ξj(s)(cid:105) = σ2δijδ(t − s). The lagged correlator is related to the zero-lagged correlator by [21]: 14 where for notational convenience we name C0 = (cid:82) xT (t)x(t)dt, i.e. the correlator at zero C0eW T t C(t) = t > 0 (A8) lag, by construction a symmetric matrix. Hence the expression for the learning update is: ∆W ∼ C0eW T t − eW tC0 S(t)dt (A9)  eWtC0 (cid:90) ∞ (cid:104) 0 t < 0 (cid:105) The temporal behavior of the weight function has been approximated by a piece-wise exponential form: S(t) =  +et/τ 0 −e−t/τ t < 0 t = 0 t > 0 (A10) where τ is STDP's time-constant, i.e. it expresses the window over which the plastic changes due to temporal coincidence are significant. Assuming that the network connections are only excitatory, and expressing without further loss of generality W = −I + A, we derive the synaptic weight update ∆W = ∆A as follows: (cid:90) ∞ ∆A = − (cid:90) ∞ 0 −t/τ eW tdt = e (cid:90) ∞ 0 0 S(t)eW tC0dt + S(t)C0eW T tdt −t/τ I+W tdt = − [W − 1/τ I] e −1 (cid:90) ∞ 0 Given that We obtain ∆A = − [W − 1/τ I] −1 C0 + C0 Leading finally to ∆A ∼ (cid:20) (cid:21)−1 I − τ 1 + τ A C0 − C0 (cid:2)W T − 1/τ I(cid:3)−1 (cid:21)−1 (cid:20) I − τ 1 + τ AT (A11) (A12) (A13) (A14) after dropping the multiplying constant τ /(1 + τ ). From this expression it is possible to derive that the weight update is anti-symmetric, and that a perfectly symmetric system would not be modified, as C0 would commute with A (see below, Eq. A20). Of course, any small initial asymmetry will eventually be blown up. We can also see that STDP's time constant also introduces the same multiplying factor τ /(1 + τ ) for A, which can be absorbed by a renormalization; we will assume therefore τ /(1 + τ ) → 1 for the remaining of the exposition. Consistently, the limiting behavior of Eq. A14 implies ∆A(τ → 0) = 0. 15 2. Minimization of loops and dynamics Now we can estimate the effect of the synaptic time-dependent plasticity expressed by Eq. A14 on the topology of the network. For this, we will postulate a penalty or energy function for what we will call "loopiness" of the network. A measure of the number of loops occurring in the network can be obtained by summing the trace of the exponentiation of the k tr (cid:2)Ak(cid:3) /k. This loop density can be simply minimized by network connectivity matrix,(cid:80) this effect; an obvious measure of the strength of the connections in a network is tr (cid:2)AAT(cid:3), making the connections vanish, so we need to introduce a regularization penalty to avoid which in a binary graph would be equivalent to the total number of links. We postulate then the following "loopiness" energy: tr (cid:2)Ak(cid:3) − 1 tr (cid:2)AAT(cid:3) E = (cid:88) k 1 k 2 The change in this energy upon small changes ∆A is expressed as ∆E ∼ tr (cid:2)∂AE∆AT(cid:3); it can be easily verified that ∂AE =(cid:0)I − AT(cid:1)−1 − A, and therefore: −1(cid:3) ∆E ∼ −tr [K1] − tr [K2] K1 = (I − AT ) −1(cid:2)(I − A) −1C0 − C0(I − AT ) (A15) K2 = A(cid:2)C0(I − AT ) −1 − (I − A) −1C0 (cid:3) (A16) We will demonstrate in what follows that the traces of K1 and K2 are strictly semi-positive under the synaptic changes elicited by STDP (i.e. Eq. A14), and therefore the loopiness energy can only decrease over time. We will consider firs that the correlation matrix can be approximated by the identity, i.e. C0 ≈ I. This case can be solved analytically; we will show numerically that the same conclusions hold for the generic case of any C0 that results from the dynamical system determined by A, drawn from a Gaussian or Poisson distribution. Let us consider tr [K1], rewritten as: tr (cid:2)(I − AT ) and which is of the from For any matrix P , −2(cid:3) −1(I − A) −1 − (I − AT ) tr (cid:2)(P P T − P 2)(cid:3) tr (cid:2)(P T − P )(P T − P )T(cid:3) = tr (cid:2)RRT(cid:3) ≥ 0 16 (A17) (A18) which upon expanding leads to tr (cid:2)P P T − P 2(cid:3) ≥ 0 (A19) and ensures the positivity of tr [K1], under the assumption C0 ≈ I. Moreover, we can show that this is valid for an arbitrary C0. Under the assumption of stability and homogeneous Gaussian noise, the correlation and weight matrices are related by the Lyapunov equation [21, 22]: W C0 + C0W T = −QQT (A20) where QQT is the generalized temperature tensor of the noise, whose components QiQj = σ2 ij are the corresponding noise variances. For the case we are considering, QQT = I, a symmetric system W = W T has the solution C0 = −W −1/2. A formal solution for the general case is [23] Assuming homogeneous noise, this reduces in our case to Following the derivation in A18-A19, the full expression for tr [K1] can be written as: (A21) (A22) (A23) (A24) (cid:3) (A25) C0 = eW T teW tdt (cid:90) ∞ (cid:90) ∞ 0 −2IteAT teAtdt e C0 = 0 (cid:3) tr [K1] = tr (cid:2)RRT C0 eA(cid:105) = tr (cid:2)(eAR)(eAR)T(cid:3) ≥ 0 Given that (cid:104) tr RRT eAT It follows through A22 that tr [K1] ≥ 0 NEW STARTS HERE @@@@@@@@ Similarly, the second term in Eq. A16, tr [K2], can be rewritten as tr (cid:2)AT (I − A) −1C0 (cid:3) − tr (cid:2)A(I − A) (cid:3) −1C0 tr (cid:2)(AT − A)(I − A) (cid:3) −1C0 which can be reduced to Replacing A(I − A)−1 by (I − A)−1 − I, the term can be transformed to tr [C0] − tr (cid:2)(I − AT )(I − A) −1C0 (cid:3) = tr [C0] − tr (cid:2)W T W −1C0 17 Assuming that QQT = I in Eq. A20, and pre-multiplying by W −1, we obtain −1C0W T = −W W −1C0] = −tr [W −1 − C0 −1] − tr [C0] tr [W T W from which we derive: tr [K2] = 2tr [C0] + tr [W −1] (A26) (A27) (A28) Now we can use the formal solution of the Lyapunov equation, Eq. A21, and the fact that for a stable matrix such as W (i.e. all its eigenvalues have negative real components) we can write −1 = − W eW tdt (cid:90) ∞ and modify Eq. A28 accordingly: tr [K2] = 2 0 tr [eW teW T t]dt − (cid:90) ∞ 0 (cid:90) ∞ 0 tr [eW t]dt (A29) Given that for any matrix, the eigenvalues satisfy λ(eW ) = eλ(W ), we get (cid:90) ∞ N(cid:88) (cid:90) ∞ N(cid:88) tr [K2] = 2 e2Re(λk)tdt − 0 k=1 0 k=1 eλktdt (A30) where λk are the N eigenvalues of W . Calling λk = −µk + iγk, µ > 0, the first term in the r.h.s. above is (cid:90) ∞ 2 0 N(cid:88) k=1 (cid:90) ∞ N(cid:88) 0 k=1 −2µktdt = e −µktdt e Now we can compare both terms in the r.h.s. of Eq. A30 for each k and t: e−µkt ≥ e−µkt+iγkt, which leads directly to tr [K2] ≥ 0 and completes the proof of the semi-negativity of the changes in the energy function (Eq. A16). Interestingly, this result is related to a property of M -matrices. For an M -matrix: (1) the off-diagonal elements are semi-negative, Mij ≤ 0 ∀i (cid:54)= j, and (2) is positive stable, Re[λi(M )] > 0 ∀i. It can be shown that for any M -matrix of dimension N (Th. 5.7.23 in [23]) tr (cid:2)M T M−1(cid:3) ≤ tr [I] = N . Choosing −W as an M -matrix, the theorem leads to a similar result for the non-negativity of tr[K2] when C0 is close to the identity. We have assumed throughout that the system is in a regime of dynamical stability, and presented a case for the stabilizing effect of STDP by linking loops and eigenvalues in Eq. 3. It follows that loopiness minimization (Eq. 2) is equivalent to maximization of stability (as 18 defined by the l.h.s. of Eq. 3), constrained by the total matrix weight. We can further effect on U = −(cid:80) understand this by explicitly expanding to first order the update equation A14 to see the i log λi. Assuming again that QQT = I, the solution to the Lyapunov equation (Eq. A21) can be approximated by a series expansion in powers of A [23]; to first approximation C0 (cid:39) I + 1 δU (cid:39) 1 more stable. 2tr [A δA + δA A], and in turn δU (cid:39)(cid:0)tr [A2] − tr (cid:2)AAT(cid:3)(cid:1) ≤ 0, making the system 2(A + AT ), leading to ∆A ∼ A − AT . Through Eq. 3 we obtain UP TO HERE @@@@ Acknowledgments We would like to acknowledge the helpful contributions of Lucas Monz´on, University of Colorado at Boulder, and Gustavo Stolovitzky, IBM Research, to the mathematical analysis. [1] Braitenberg V. and Schuz A. Cortex: Statistics and geometry of neuronal connectivity (2nd ed.). Berlin: Springer-Verlag, 1998. [2] Stepanyants A., Hirsch J.A., Martinez L.M., Kisvrday Z.F., Ferecsk A.S., and Chklovskii D.B. Local potential connectivity in cat primary visual cortex. Cerebral Cortex, 18(1):13 -- 28, 2007. [3] Crick F. and Koch C. Constraints on cortical and thalamic projections: the no-strong-loops hypothesis. Nature, 391:245 -- 250, 1998. [4] Le Be J-V. and Markram H. Spontaneous and evoked synaptic rewiring in the neonatal neocortex. Proc. Natl. Acad. Sci. USA, 103:13214 -- 13219, 2006. [5] Markram H., Lubke J., Frotscher M., and Sakmann B. Regulation of synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science, 275:213 -- 215, 1997. [6] Morrison A., Diesmann M., and Gerstner W. Phenomenological models of synaptic plasticity based on spike timing. Biol. Cybern., 98:459 -- 478, 2008. [7] Abbott L.F. and Nelson S. B. Synaptic plasticity: taming the beast. Nature Neuroscience, 3(11s):1178 -- 1183, 2000. [8] Song S., Miller K.D., and Abbott L.F. Competitive hebbian learning through spike-timing- dependent synaptic plasticity. Nat. Neurosci., 3:919 -- 926, 2000. 19 [9] Burkitt A.N., Meffin H., and Grayden D.B. Spike-timing-dependent plasticity: The relation- ship to rate-based learning for models with weight dynamics determined by a stable fixed point. Neural Comput., 16:885 -- 940, 2004. [10] Gutig R., Aharonov R., Rotter S., and Sompolinsky H. Learning input correlations through nonlinear temporally asymmetric hebbian plasticity. J. Neurosci., 23:3697 -- 3714, 2003. [11] Ma'ayan A., Cecchi G.A., Wagner J., Rao A.R., Iyengar R., and Stolovitzky G. Ordered cyclic motifs contributes to dynamic stability in biological and engineered networks. Proc. Natl. Acad. Sci. USA, 105(49):19235 -- 40, 2008. [12] Fino E., Deniau J-M., and Venance L. Cell-specific spike-timing-dependent plasticity in gabaergic and cholinergic interneurons in corticostriatal rat brain slices. J. Physiol., 586(1):265 -- 282, 2008. [13] Seol G.H., Ziburkus J., Huang S., Song L., Kim I.T., and Takamiya K. Neuromodulators control the polarity of spike-timing-dependent synaptic plasticity. Cell, 55:919 -- 929, 2007. [14] Prasolov V.V. Problems and theorems in linear algebra. American Mathematical Society, 1994. [15] Song S., Sjostrom P.J., Reigl M., Nelson S., and Chklovskii D.B. Highly nonrandom features of synaptic connectivity in local cortical circuits. PLoS biology, 3(3):507 -- 509, 2005. [16] Fino E., Glowinski J., and Venance L. Bidirectional activity-dependent plasticity at cortico- striatal synapses. J. Neuroscience, 25(49):11279 -- 11287, 2005. [17] Abeles M. Corticonics: Neural Circuits of the Cerebral Cortex. Cambridge, 1991. [18] Hosaka R., Araki O., and Ikeguchi T. Stdp provides the substrate for igniting synfire chains by spatiotemporal input patterns. Neural Computation, 20:415 -- 435, 2008. [19] Pawlak V. and Kerr J.N.D. Dopamine receptor activation is required for corticostriatal spike- timing dependent plasticity. J. Neuroscience, 28(10):2435 -- 2446, 2008. [20] Thivierge J.P. and Cisek P. Nonperiodic synchronization in heterogeneous networks of spiking neurons. J. Neuroscience, 28(32):7968 -- 7978, 2008. [21] Risken H. The Fokker Planck Equation: Methods of Solutions and Applications (2nd Ed.). Springer, 1996. [22] DelSole T. Stochastic models of shear-flow turbulence with enstrophy transfer to subgrid scales. J. Atmosph. Sci., 56:3692 -- 3703, 1999. [23] Horn R.A. and Johnson C.R.J. Topics in Matrix Analysis. Cambridge University Press, 1991. 20 [24] We use microcircuitry to refer to neural circuitry observed at the level of neurons and synapses and not necessarily in reference to a restricted spatial extent of these neurons (e.g., a "col- umn"). For this reason, we view every brain connection as part of some microcircuit topology. [25] Note that paths that traverse the same network node more than once are also counted. 21 FIG. 1: Schematic of the topological effect of STDP. Feedback connections in an initial topology (left) from first (1) and second (2) order "follower" neurons (light blue) to a "trigger" neuron (green) create loops of length k = 2 and k = 3. These connections are selectively penalized by the STDP learning rule (lower middle, red). The plot (middle) depicts this rule, with the time difference between follower (black) and trigger (green) action potentials on the x-axis, and the expected synaptic modification on the y-axis. When spikes successfully propagate through the loopy network they feed back to the trigger, arriving at the follower-trigger synapse immediately after the trigger neuron fired and resulting in synaptic depression (red). Through repeated spike propagation events, STDP results in a completely feed forward learned topology (right) to the output neuron (dark blue). 22 121212INITIAL TOPOLOGYLEARNED TOPOLOGYLOOP-REGULATING STDP∆t∆w FIG. 2: Global topological effects of STDP. A) A monotonic decrease in the loopiness measure (see Equation 2, first term) over time is observed in a simulated network of 100 neurons undergoing STDP. Simultaneously, STDP results in a net increase in the weightedness of the network (inset, see Equation 2, second term). Shown here and in B) is the average of 8 separate simulations of 20 seconds of network activity; error bars are standard deviation. B) Number of closed loops of length 5, 3, and 2, decreases as a function of weight threshold for network connections. Dotted lines show counts for randomized networks with same number of total connections. C) Number of unique loops sampled from five networks with varying synaptic delays following 10 seconds of simulated activity, and from a random network with 5, 000 connections. Number of loops is shown as a function of loop length. Loops were sampled across different learned networks while maintaining the number of network connections at 5, 000 by varying the weight threshold (from 0.003 to 0.0046) for each delay. Greater synaptic delays decreases the loop-eliminating topological effect of standard STDP. 23 00.0020.0040.0060.0080.011001051010Adjacency weight thresholdNumber of closed loopsFewer Closed Loops Than Randomlen=5len=3len=2 randomSTDP051015202510-2100102104106Fewer Unique Loops, Varying DelayLoop lengthNumber of unique loops random4.0 ms2.01.00.50.1051015200.050.10.150.2Time (secs)Decrease In Network Loopiness100n=11ktr(AT)k100n=112trAAT05101520121314151617181920Time (secs)Increase In Network WeightednessABC FIG. 3: Local topological effects of STDP. A) Inverse relationship of in-degree versus out-degree of intra-network connections for each neuron in a network after 10 seconds of STDP across multiple weight thresholds for network connections. Colors in left and right panels correspond to a weight threshold used to construct the network over which degrees were measured. The color key can be read from the left panels' vertical axes and corresponding color found along each manifold. B) Correlated extra-network in-degree and intra-network out-degree indicate an opposite effect of STDP on extra-network inputs. 24 05010005010000.0050.010.015koutkinAdjacency weight threshold020040060005010000.0050.010.015koutkinAdjacency weight threshold020406080100020406080100kinkoutIntra-network In-Degree, Intra-network Out-Degree0100200300400500020406080100kinkoutExtra-network In-Degree, Intra-network Out-DegreeAB FIG. 4: Effect of reverse STDP. A) In-degree versus out-degree of intra-network connections fol- lowing different durations and polarities of STDP, shows a strong inverse relationship for standard STDP. Adjacency weight threshold was 0.005. Green markers correspond to the network after 1 second of standard STDP, followed by 3 or 5 seconds of standard STDP (blue markers) or reverse STDP (red markers). B) Total synaptic input versus output for intra-network connections shows a similar inverse relationship for standard STDP. C) Number of closed loops of length 5, 3, and 2, is decreased by standard STDP, and restored with reverse STDP (plotted as in Fig. 2B). D) In-degree of extra-network inputs versus out-degree of intra-network outputs, plotted as in A, shows a strong positive correlation for standard STDP. Adjacency weight threshold was 0.007 for extra-network inputs and 0.005 for intra-network outputs. E) Total synaptic extra-network input versus total synaptic intra-network output, plotted as in B, shows a similar positive correlation. In each panel the topological effects of STDP are reversed by reverse STDP. 25 00.20.40.60.810.10.20.30.40.50.60.70.80.91Σ winΣ wout0204060801000102030405060708090100kinkoutIntra-network Inputs, Intra-network Outputs 2.830.10.20.30.40.50.60.70.80.912002500102030405060708090100Extra-netw1:STDP 1 s2:STDP 3 s3:STDP 5 s2:RSTDP 3 s3:RSTDP 5 s00.0050.011001051010len=5len=3len=200.0050.011001051010len=5len=3len=200.0050.011001051010Number of closed loopslen=5len=3len=200.0050.011001051010Adjacency weight thresholdlen=5len=3len=2ADCΣ woutkout1234BE FIG. 5: Dynamical effects of STDP. A) Raster plot of the spiking activity for a network after STDP (top), and for a surrogate network where intra-network weights were reassigned randomly to network connections, thus destroying STDP-learned topology (bottom). Each point corresponds to a spike for each neuron. Each neuron was assigned a rank according to the sum of its extra- network input weights, with the lowest rank corresponding to the highest sum. B) Peri-event time histograms for each neuron in the STDP network (top left) and its surrogate (bottom left), pooled across 8 separate simulations (bin width, 2msec). Histograms show different network propaga- tion properties. Spike counts and extra-network weights for the same networks do not co-vary in the STDP-learned topology (top right), but are highly correlated for the surrogate (bottom right). C) Peri-stimulus time histograms summed across all simulations and all neurons for the STDP network (blue) and three surrogates, in which the intra-network connections (red), extra- network connections (green) or both (magenta) were randomized (top). Skewness versus kurtosis 26 of these histograms averaged across 8 separate simulations each (bottom, error bars show stan- 24682.62.833.23.4Spikes/secΣ win, extranet0510152.62.833.23.4Spikes/secΣ win, extranet1010.51111.51212.51302040608010020406080100Peri-event time (msec)Neuron Rank, Σ win, extranetNon-randomized Network1010.51111.51212.51302040608010020406080100Peri-event time (msec)Neuron Rank, Σ win, extranetIntra-network-randomized Network-15-10-505101520-20-15-10-505101520-20ABt (sec)t (sec)Non-randomized NetworkIntra-network-randomized Network
1605.05291
1
1605
2016-05-17T19:14:03
Multistep Model for Predicting Upper-Limb 3D Isometric Force Application from Pre-Movement Electrocorticographic Features
[ "q-bio.NC", "cs.HC" ]
Neural correlates of movement planning onset and direction may be present in human electrocorticography in the signal dynamics of both motor and non-motor cortical regions. We use a three-stage model of jPCA reduced-rank hidden Markov model (jPCA-RR-HMM), regularized shrunken-centroid discriminant analysis (RDA), and LASSO regression to extract direction-sensitive planning information and movement onset in an upper-limb 3D isometric force task in a human subject. This mode achieves a relatively high true positive force-onset prediction rate of 60% within 250ms, and an above-chance 36% accuracy (17% chance) in predicting one of six planned 3D directions of isometric force using pre-movement signals. We also find direction-distinguishing information up to 400ms before force onset in the pre-movement signals, captured by electrodes placed over the limb-ipsilateral dorsal premotor regions. This approach can contribute to more accurate decoding of higher-level movement goals, at earlier timescales, and inform sensor placement. Our results also contribute to further understanding of the spatiotemporal features of human motor planning.
q-bio.NC
q-bio
Multistep Model for Predicting Upper-Limb 3D Isometric Force Application from Pre-Movement Electrocorticographic Features* Jing Wu, Student Member, IEEE, Benjamin R. Shuman, Bingni W. Brunton, Katherine M. Steele, Jared D. Olson, Rajesh P.N. Rao, Member, IEEE, and Jeffrey G. Ojemann  Abstract-Neural correlates of movement planning onset and direction may be present in human electrocorticography in the signal dynamics of both motor and non-motor cortical regions. We use a three-stage model of jPCA reduced-rank hidden Markov model (jPCA-RR-HMM), regularized shrunken- centroid discriminant analysis (RDA), and LASSO regression to extract direction-sensitive planning information and movement onset in an upper-limb 3D isometric force task in a human subject. This mode achieves a relatively high true positive force-onset prediction rate of 60% within 250ms, and an above-chance 36% accuracy (17% chance) in predicting one of six planned 3D directions of isometric force using pre- movement signals. We also find direction-distinguishing information up to 400ms before force onset in the pre- movement signals, captured by electrodes placed over the limb- ipsilateral dorsal premotor regions. This approach can contribute to more accurate decoding of higher-level movement goals, at earlier timescales, and inform sensor placement. Our results also contribute to further understanding of the spatiotemporal features of human motor planning. I. INTRODUCTION Decoding human arm movement has been a major goal of motor brain-computer interfaces (BCIs) for the development upper-limb neuroprostheses. Many previous decoding approaches have related recorded sensorimotor signals with specific virtual or robotic movements [1]–[3], allowing a potential BCI to reconstruct intended movements from a combination of basic signal changes. For example, primary motor cortex signals during individual finger movements could be combined to reconstruct multi-axis virtual cursor movements [4], [5], or pinch or grasp movements [6]–[8]. Alternatively, more of movements, or movement plans, may be present in recorded representations abstract *Research supported by grants from NSF EEC-1028725, NINDS 5R01NS065186, NIH 2K12HD001097, NIH 5U10NS086525, and the WRF Fund for Innovation in Neuroengineering. J. Wu is with the department of Bioengineering, University of Washington, Seattle, WA 98105 USA. (phone: 858-729-8669; e-mail: [email protected]). B.R. Shuman is with the department of Mechanical Engineering, (e-mail: University of Washington, Seattle, WA 98105 USA [email protected]) B.W. Brunton is with the department of Biology, University of Washington, Seattle, WA 98105 USA. (e-mail: [email protected]). K.M. Steele is with the department of Mechanical Engineering, (email: University of Washington, Seattle, WA 98105 USA [email protected]) J.D. Olson is with the department of Rehabilitation Medicine, University of Washington, Seattle, WA 98105 USA. (email: [email protected]) R.P.N. Rao is with the department of Computer Science, University of Washington, Seattle, WA 98105 USA. (e-mail: [email protected]). J.G. Ojemann is with the department of Neurosurgery, University of Washington, Seattle, WA 98105 USA. (e-mail: [email protected]). brain signals prior to the onset of movement. The presence of these signals may especially be true in areas outside of the sensorimotor cortex. For instance, studies in non-human primates and human neuroimaging have established that, during an upper-extremity movement in a reach task [9]– [11], the parietal and premotor cortices play a key role in the representation of target trajectories and executing planned movements in the desired spatial coordinates. implanted for medically An effective control signal for a BCI must maximize the information available, including preparation prior to the onset of movement. Patients with indwelling electrocorticography (ECoG) electrodes intractable epilepsy provide recordings with centimeter resolution combined with millisecond sampling over a large brain region. These signals offer an opportunity to detect correlates of the entire movement preparation and execution process. Previous work have demonstrated the capacity of ECoG signals to resolve the timing and interaction between regions during finger movements [8], upper-limb movements [1], [2], as well as premotor responses that precede the sensory and motor cortex activation up to 100ms in advance [12]. One previous work has also shown some success in detecting ECoG representations of onset and direction of hand-arm movements [13]. However, the temporal and spatial distributions of the recorded signal in motor preparatory regions are not well- established in human ECoG. It is yet unclear which non- motor regions contain ECoG signals with direction- distinguishing preparatory information, whether these signals are localized to just the contralateral hemisphere or both hemispheres, and how they are localized in time. It is also unknown whether correlates of preparatory activity are limited to reach trajectories, or also can generalize to other directional tasks such as isometric force application towards planned directions. These unknown factors are particularly relevant to the engineering of BCI, as signals conveying information about the planning of a movement may allow for better resolution of differing task goals, inform locations for sensor placement, and enable decoding of these goals earlier than signals in the sensorimotor cortex. In this work, we investigated the recorded ECoG activities of a patient with fortuitous bilateral implantation of electrodes in the parietal and arm-associated sensorimotor cortices. The patient performed a right arm isometric force task toward designated directions. We constructed a three- stage model to examine the localization and timing of direction-discriminable information, allowing for the parallel prediction of isometric force onset and direction. II. METHODS A. Experimental Task and Data Acquisition The patient is a 26-year-old right-handed male with intractable seizures thought to be related to a medial parietal lesion; invasive monitoring of bilateral parietal regions was recommended. The subject received pre-operative anatomical MRI, and was implanted with two frontoparietal 2×8 platinum electrode grids and two 1×8 strips (Ad-Tech, Racine, WA, USA), with 3mm electrode-diameters embedded in silastic with center-to-center distances of 10mm. Subsequently, X-ray CT imaging captured the location of the electrodes. These electrodes were sampled at 1220.7 Hz on a 128-channel Tucker-Davis Technology (Alachua, FL, USA) System 3 RZ5D Neurophysiology Base Processor with associated PZ5 NeuroDigitizer. The subject had normal motor function pre-operatively and throughout the experiments and inpatient stay. The patient gave written, informed consent through a protocol approved by the University of Washington Institutional Review Board. The patient was comfortably positioned in front of one screen during the task, and was visually cued to apply isometric force in one of 6 directions (up, down, left, right, forward, back) to an affixed AMTI force transducer handle (Watertown, MA, USA) held in the right hand. Each trial consisted of a hold with approximately 15 lbs (67 N) of force and 2 seconds in duration, with an inter-trial interval of 3 seconds. 10 trials of each of 6 directions were recorded for a total of 60 isometric force trials. B. Data Preprocessing and Feature Extraction using Statistical Parametric Pre-operative MRI was co-registered with postoperative Mapping CT (http://www.fil.ion.ucl.ac.uk/spm), with the pial surface reconstructed with FreeSurfer (http://freesurfer.net) and custom mapping and projection code [14] implemented in MATLAB (Natick, MA, USA). The two frontoparietal hemisphere 2×8 grid recordings were visually inspected for data quality, and channels with high level of artifacts and low impedance were rejected (1 in the contralateral L grid and 2 in R). The remaining 29 recorded channels were filtered with a common average reference filter siʹ=si−J−1∑jsj for J non- rejected channels within each separate lateral grid. The filtered signals were then analyzed for time-frequency content by continuous wavelet transform with the non- analytic Morlet wavelet defined by Ψ(sω) = π−¼e[−(sω−ω0)^2]/2, where ω0=6 and s reflects log scaling factors to obtain ⅛- octave resolution across pseudo-frequencies 2−200 Hz. We used the absolute magnitude of the wavelet coefficient time series at each scaling, and also extracted the lowpassed ECoG local motor potential (LMP) for each channel as outlined in [15] as an additional time series. All time series were binned by 50ms and normalized to unit variance. Recorded force transducer voltages were converted to force using manufacturer sensitivity calibration data; Cartesian forces were transformed to spherical coordinates to obtain task force direction and magnitude. Subsequently, the initiation and end of recorded isometric force in each trial were carefully hand-labeled (examples shown with dotted force traces and vertical onset lines in Figure 3). C. Pre-Movement Classification of Force Direction jPCA-RR-HMM analysis: We constructed a reduced-rank hidden Markov model (RR-HMM) [16] from the normalized binned time series using jPCA, a dimensionality reduction method designed to find orthonormal bases that also exhibit a rotational structure, with the expectation of a rhythmic repeating behavior from the signal dynamics [17]. To find these repeating mean trajectory templates, filtered time series were split into 2-second segments center-aligned to the initiation of movement, reflecting a time period of [−1000ms…+1000ms] for each movement trial. The high- dimensional time series for each channel × frequency xw,t,c for [t-1000, t-950, …t0…t+1000] were averaged for each of the 6 movement direction conditions d = [1…6]. jPCA was used on these mean timeseries to obtain a set of jPC coefficients that describe a 10-dimensional space to maximize rotational dynamics in these mean ECoG signal trajectories. The original trial segments xw,t,c,d for [t-1000…t+1000] were then projected through these jPCs, such that each trial segment is now reduced to a 10D space as xʹ[jPC1,2],t,d for [t-1000…t+1000]. All time points were then labeled as belonging to one of 12 clusters agnostic to condition, using unsupervised hierarchical clustering of all xʹ time points with Ward's method, which recursively locates clusters that minimize within-cluster distance variance in the dimensionally-reduced Euclidean space. This method generated a list of observed states st,d for t = [−1000ms … +1000ms], where all observed s are categorical class labels, s ϵ [1…12], and d is one of six directions of movement. The pre-movement cluster-time-series were then isolated and used as the sole training data to construct the HMM. Cluster labels immediately preceding movement st,d for t = [−500ms…0ms] were grouped by movement direction condition, and used as observed sequences to train six hidden Markov models Hd, one for each force movement direction with d = [1…6]. The Baum-Welch algorithm was used with 12 observed states and 8 hidden states to train each Hd. Direction predictions were made with leave-one-out validation, where each HMM generated from training st,d was used to calculate the posterior state probability of observing the sequences in the testing st,d. The direction associated with the Hd giving the highest posterior probability (smallest negative log-probability) was used as the classified direction. Regularized discriminant analysis: The normalized, binned time series were split into 1-second chunks aligned to the initiation of movement segments, reflecting a time period of the start corresponding to 1 second before movement initiation. Each feature xw,t,c at wavelet scale w, timepoint t, and channel c is used as a feature for the training of a multiclass regularized discriminant analysis model with 10-fold validation, where the posterior probability that observation x belongs to class k follows [−1000ms…0ms] trial, with for each Shrunken centroids regularization as described in [18] attempts to reduce overfitting bias by (a) introducing a modified covariance matrix to stabilize the cross-validated sample covariance Σ and (b) applying a threshold to a modified correlation matrix to prune down the number of useful predictors, as described by where, This is accomplished by iteratively searching through parameters γ and δ, where γ modifies the sample covariance matrix Σ, and δ thresholds the correlation matrix C and reduces the number of predictors by thresholding them from the posterior probability evaluation [18]. D. Movement Onset and Force Magnitude Prediction (Lasso) regression The normalized binned time series was fit to the magnitude of the force using least absolute shrinkage and selection operator [19]. Two-fold validation was used wherein the whole time series of the each half of the recorded movement session rw,c, with channel × frequency wavelet coefficients through time was used to fit linear regression coefficients β where predicted ŷt = βw,c,t rw,c,t + β0. This generates βw,c,t as a function of a regularization parameter λ, which thresholds the number of nonzero β coefficients. λ was chosen such that the mean-squared-error is the lowest in validation, which resulted in 132 nonzero β coefficients, which were used to predict the forces in the other half of the recorded session. The predicted force time series ŷt were then smoothed using a 1st-order 2 Hz, −10dB Butterworth lowpass filter and numerically differentiated to obtain dŷt/dt. Peak analysis was used to find the local maxima of dŷt/dt using a fixed threshold, and the local maxima points represent potential times of force onset. III. RESULTS A. Force Direction Classification Using jPCA-RR-HMM An average HMM classification accuracy of 35.6% was obtained using pre-movement clustered timepoints 500ms prior to the onset of movement. Accuracies were computed using force application directions using jPCA-projected, hierarchically clustered pre-movement ECoG. This performance is signifi- leave-one-out cross-validation between 6 Figure 1. Confusion matrix of the model classification accuracies using jPCA-RR-HMM, in 59 valid isometric force application trials in 6 directions using pre-movement ECoG recordings. Regularized features in time and space 107 7 Ch 26 Ch 5 Ch 17 -400 -300 -200 -100 200 180 160 140 120 100 80 60 40 20 ) z H ( y c n e u q e r f o d u e s p t l e e v a W -500 ) s s e l t i n u ( l e u a v 6 5 4 3 2 1 0 0 Time before movement (ms) Figure 2. Left: Blue = recorded electrodes; Yellow = recorded electrodes corresponding to regularized features in Right plot; Gray = rejected for artifacts. Right: Time-space plot of salient features from regularization. Each block refers to one feature, most salient (high δ) features in yellow. cantly higher than chance of 17%. Figure 1 shows the confusion matrix of classification in the 6 directions, where the main diagonal shows the correct classifications. Five of six directions showed accuracies above chance, with the most accurate direction being isometric movement back toward the torso, which also had very low false-prediction rates. B. Time-Space Localization of Directional Information In a second analysis, we used regularized discriminant analysis directly on the wavelet features of recorded pre- movement ECoG signals to obtain a cross-validated directional classification accuracy of 27%. While this is not as high as the outcome obtained with the jPCA-RR-HMM analysis, RDA regularization and shrinkage operates on the original feature space, and the remaining 27 features included in computing the posterior class probability must exceed the regularization threshold set by δ as outlined in equation (5). These remaining wavelet coefficient features demonstrably contain directionally-discriminable information, and are time and space. Interestingly, the spatiotemporal distribution of these features as shown in Fig. 2 reveal that the earliest information about movement is present in channel 26 (dorsal premotor cortex), then passes on to channel 17, and finally to channel 5 (primary motor cortex). C. Lasso Regression Onset and Force Prediction Results cross-validated localized in Cross-validated force predictions agree with overall force magnitude (R2 = 0.42), but a strong predictive relationship between peak force magnitude and ECoG signals magnitude in this task was not found. However, the smoothed predicted e c r o f N Figure 3. Force and onset timing predictions from Lasso regression. Triangles: predicted force onset; dotted vertical lines: actual onset. force time-series can be numerically derived to obtain an accurate assessment of force onset, predicting onset to within 250ms at a 60% accuracy and within 500ms at an 80% accuracy. The false positive predictive rate was 22%. IV. DISCUSSION A. Direction-Discriminating Dynamics We showed that the combination of dimensionality reduction and HMM allowed for RR-HMM states to capture complex multidimensional dynamics in a relatively small number of observed states. The performance of this technique suggests that the dynamics between ECoG timepoints during movement preparatory behavior may carry useful task- relevant information. Though the predictive performance of this technique is not sufficient alone to drive brain-computer interfaces, it may be used in combination with techniques such as Kalman filters that can benefit from earlier probability updates to improve the inference of user intent. B. Spatiotemporal Feature Localization the showed We found that wavelet-based spectral characteristics of ECoG recordings can discriminate between different directions of isometric arm movements significantly better than chance in a regularized discriminant analysis, despite the overlap between the recruited muscle groups and the isometric task. We find that signals carrying direction- discriminatory information occurred over the ipsilateral premotor cortex up to 400ms before the onset of detected force, in contrast to the signals from the contralateral sensorimotor strongest discrimination ability starting roughly 100−150msec prior to movement onset (Fig. 2). The timing of the information in ipsilateral changes is particular interesting in potentially offering a technique to screen for neural processes spanning multiple cortical regions, and may give hints to the possible timing of distributed trajectory planning processes. C. Differences Between Movement Trajectory and cortex which Magnitude While the ECoG signal can be used to both decode direction and force onset in parallel, we found in this dataset that high-performing regression features in predicting force changes do not also perform well in discriminating between directions, potentially suggesting separate usable ECoG signatures for motor execution and planning. D. Future Considerations Further work is necessary to determine whether these particular correlates of movement planning generalize across tasks to other types of upper-limb movements, or generalizes to other measures of force application such as non-task muscle contraction. We also wish to examine the possibility of learning effects with repeated practice. ACKNOWLEDGMENT We would like to thank Lila Levinson for her outstanding exploratory analyses on this dataset. [4] [5] [6] [7] [8] [9] [10] [11] [13] REFERENCES [1] T. Pistohl, T. Ball, A. Schulze-Bonhage, A. Aertsen, and C. Mehring, "Prediction of arm movement trajectories from ECoG- recordings in humans," J. Neurosci. Methods, vol. 167, no. 1, pp. 105–114, Jan. 2008. N. R. Anderson, T. Blakely, G. Schalk, E. C. Leuthardt, and D. W. Moran, "Electrocorticographic (ECoG) correlates of human arm movements," Exp. Brain Res., vol. 223, no. 1, pp. 1–10, Sep. 2012. [3] M. S. Fifer, S. Acharya, H. L. Benz, M. Mollazadeh, N. E. Crone, [2] and N. V. Thakor, "Towards Electrocorticographic Control of a Dexterous Upper Limb Prosthesis," IEEE Pulse, vol. 3, no. 1, pp. 38–42, Jan. 2012. J. R. Wolpaw, N. Birbaumer, D. J. McFarland, G. Pfurtscheller, and T. M. Vaughan, "Brain–computer interfaces for communication and control," Clin. Neurophysiol., vol. 113, no. 6, pp. 767–791, Jun. 2002. G. Schalk, K. J. Miller, N. R. Anderson, J. A. Wilson, M. D. Smyth, J. G. Ojemann, D. W. Moran, J. R. Wolpaw, and E. C. Leuthardt, "Two-dimensional movement control using electrocorticographic signals in humans," J. Neural Eng., vol. 5, no. 1, p. 75, Mar. 2008. P. Shenoy, K. J. Miller, J. G. Ojemann, and R. P. N. Rao, "Finger Movement Classification for an Electrocorticographic BCI," in 3rd International IEEE/EMBS Conference on Neural Engineering, 2007. CNE '07, 2007, pp. 192–195. J. Kubánek, K. J. Miller, J. G. Ojemann, J. R. Wolpaw, and G. Schalk, "Decoding flexion of individual fingers using electrocorticographic signals in humans," J. Neural Eng., vol. 6, no. 6, p. 066001, Dec. 2009. K. J. Miller, C. J. Honey, D. Hermes, R. P. Rao, M. denNijs, and J. G. Ojemann, "Broadband changes in the cortical surface potential track activation of functionally diverse neuronal populations," NeuroImage, vol. 85, Part 2, pp. 711–720, Jan. 2014. L. R. Bremner and R. A. Andersen, "Coding of the Reach Vector in Parietal Area 5d," Neuron, vol. 75, no. 2, pp. 342–351, Jul. 2012. P. S. Archambault, S. Ferrari-Toniolo, R. Caminiti, and A. Battaglia-Mayer, "Visually-guided correction of hand reaching movements: The neurophysiological bases in the cerebral cortex," Vision Res., Sep. 2014. S. Schaffelhofer, A. Agudelo-Toro, and H. Scherberger, "Decoding a Wide Range of Hand Configurations from Macaque Motor, Premotor, and Parietal Cortices," J. Neurosci., vol. 35, no. 3, pp. 1068–1081, Jan. 2015. [12] H. Sun, T. M. Blakely, F. Darvas, J. D. Wander, L. A. Johnson, D. K. Su, K. J. Miller, E. E. Fetz, and J. G. Ojemann, "Sequential activation of premotor, primary somatosensory and primary motor areas in humans during cued finger movements," Clin. Neurophysiol., vol. 126, no. 11, pp. 2150–2161, Nov. 2015. Z. Wang, A. Gunduz, P. Brunner, A. L. Ritaccio, Q. Ji, and G. Schalk, "Decoding onset and direction of movements using Electrocorticographic (ECoG) signals in humans," Front. Neuroengineering, vol. 5, p. 15, 2012. [14] D. Hermes, K. J. Miller, H. J. Noordmans, M. J. Vansteensel, and N. F. Ramsey, "Automated electrocorticographic electrode localization on individually rendered brain surfaces," J. Neurosci. Methods, vol. 185, no. 2, pp. 293–298, Jan. 2010. [15] G. Schalk, J. Kubánek, K. J. Miller, N. R. Anderson, E. C. Leuthardt, J. G. Ojemann, D. Limbrick, D. Moran, L. A. Gerhardt, and J. R. Wolpaw, "Decoding two-dimensional movement trajectories using electrocorticographic signals in humans," J. Neural Eng., vol. 4, no. 3, p. 264, Sep. 2007. S. M. Siddiqi, B. Boots, and G. J. Gordon, "Reduced-Rank Hidden Markov Models," ArXiv09100902 Cs, Oct. 2009. [16] [17] M. M. Churchland, J. P. Cunningham, M. T. Kaufman, J. D. Foster, P. Nuyujukian, S. I. Ryu, and K. V. Shenoy, "Neural population dynamics during reaching," Nature, vol. 487, no. 7405, pp. 51–56, Jul. 2012. [18] Y. Guo, T. Hastie, and R. Tibshirani, "Regularized linear discriminant analysis and its application in microarrays," Biostatistics, vol. 8, no. 1, pp. 86–100, Jan. 2007. [19] R. Tibshirani, "Regression Shrinkage and Selection via the Lasso," J. R. Stat. Soc. Ser. B Methodol., vol. 58, no. 1, pp. 267–288, 1996.
1012.3623
1
1012
2010-12-16T14:37:47
Information capacity and transmission are maximized in balanced cortical networks with neuronal avalanches
[ "q-bio.NC", "cond-mat.dis-nn", "physics.bio-ph" ]
The repertoire of neural activity patterns that a cortical network can produce constrains the network's ability to transfer and process information. Here, we measured activity patterns obtained from multi-site local field potential (LFP) recordings in cortex cultures, urethane anesthetized rats, and awake macaque monkeys. First, we quantified the information capacity of the pattern repertoire of ongoing and stimulus-evoked activity using Shannon entropy. Next, we quantified the efficacy of information transmission between stimulus and response using mutual information. By systematically changing the ratio of excitation/inhibition (E/I) in vitro and in a network model, we discovered that both information capacity and information transmission are maximized at a particular intermediate E/I, at which ongoing activity emerges as neuronal avalanches. Next, we used our in vitro and model results to correctly predict in vivo information capacity and interactions between neuronal groups during ongoing activity. Close agreement between our experiments and model suggest that neuronal avalanches and peak information capacity arise due to criticality and are general properties of cortical networks with balanced E/I.
q-bio.NC
q-bio
Information capacity and transmission are maximized in balanced cortical networks with neuronal avalanches Woodrow L. Shew*1, Hongdian Yang (杨黉典)*1,2, Shan Yu (余山)1, Rajarshi Roy2, Dietmar Plenz1 1 Section on Critical Brain Dynamics, Laboratory of Systems Neuroscience, National Institutes of Mental Health, Bethesda, MD, USA 20892; 2 Institute for Physical Science and Technology, University of Maryland, College Park, MD, USA 20742 * These authors contributed equally to this work The repertoire of neural activity patterns that a cortical network can produce constrains the network’s ability to transfer and process information. Here, we measured activity patterns obtained from multi-site local field potential (LFP) recordings in cortex cultures, urethane anesthetized rats, and awake macaque monkeys. First, we quantified the information capacity of the pattern repertoire of ongoing and stimulus- evoked activity using Shannon entropy. Next, we quantified the efficacy of information transmission between stimulus and response using mutual information. By systematically changing the ratio of excitation/inhibition (E/I) in vitro and in a network model, we discovered that both information capacity and information transmission are maximized at a particular intermediate E/I, at which ongoing activity emerges as neuronal avalanches. Next, we used our in vitro and model results to correctly predict in vivo information capacity and interactions between neuronal groups during ongoing activity. Close agreement between our experiments and model suggest that neuronal avalanches and peak information capacity arise due to criticality and are general properties of cortical networks with balanced E/I. 2 INTRODUCTION In the cortex, populations of neurons continuously receive input from upstream neurons, integrate it with their own ongoing activity, and generate output destined for downstream neurons. Such cortical information processing and transmission is limited by the repertoire of different activated configurations available to the population. The extent of this repertoire may be quantified by its entropy H; in the context of information theory, entropy characterizes the information capacity of the population (Shannon, 1948; Rieke et al., 1997; Dayan and Abbott, 2001). Information capacity is important because, as the name suggests, it defines upper limits on aspects of information processing. For example, consider the information transmitted from input to output by a population that only has two states in its repertoire (H≤1 bit). No matter how much information the input contains, the information content of its output cannot exceed 1 bit. A network with low entropy presents a bottleneck for information transmission in the cortex. Thus, it is important to understand the mechanisms that modulate the entropy of cortical networks. Cortical activity depends on the ratio of fast excitatory (E) to inhibitory (I) synaptic inputs to neurons in the network. This E/I ratio remains fixed on average even during highly fluctuating activity levels (Shu et al., 2003; Okun and Lampl, 2008). However, it is currently not known whether a particular E/I ratio is advantageous for certain aspects of information processing. The existence of such an optimal ratio is suggested by two competing effects of E/I on entropy. First, a large E/I ratio, i.e. if excitation is insufficiently restrained by inhibition, can cause very high correlations between neurons (Dichter and Ayala, 1987). Since increased correlations decrease entropy (Rieke et al., 1997; Dayan and Abbot, 2001), we anticipate that a sufficiently large E/I ratio limits information transmission. This is consistent with findings that moderate levels of correlation can play an important role in population coding (Pola et al., 2003; Jacobs et al., 2009). At the other extreme, i.e. a small E/I ratio, weak excitatory drive is expected to reduce correlations as well as the overall level of neural activity. Although reduced correlations can lead to higher entropy, this increase may be counteracted by a concurrent drop in activity. Sufficiently suppressed activity reduces entropy (Rieke et al., 1997; Dayan and Abbot, 2001). Accordingly, we hypothesize that cortical entropy and information transmission are maximized for an intermediate E/I ratio. Here we tested our hypothesis experimentally in cortex cultures, anesthetized rats, and awake monkeys and compared our results with predictions from a computational model. We discovered an optimal intermediate E/I ratio distinguished by 1) maximal entropy and 2) maximal information transmission between input and network output. This finding was based on analysis of both ongoing and stimulus-evoked population activity. Moreover, at this optimal E/I ratio, ongoing activity emerges in the form of neuronal avalanches (Beggs and Plenz, 2003) and interactions within the network are moderate. Agreement with our model suggests that by maintaining this particular E/I ratio the cortex operates near criticality and optimizes information processing. MATERIAL AND METHODS 3 In vitro MEA recordings and pharmacology. Cortex-VTA (ventral tegmental area) organotypic co-cultures were grown on planar integrated multi-electrode arrays (MEA; for details see Shew et al., 2009). In brief, the MEA contained 60 recording electrodes (8x8 grid with no corner electrodes, 30 μm diameter, 200 μm inter-electrode spacing). LFP signals were sampled at 4 kHz and low-pass filtered (50 Hz cut off). Experiments consisted of a 1 hour recording of ongoing activity, followed by 0.5 – 1 hr of stimulus- evoked activity. Next, the antagonists were bath-applied and recordings of ongoing activity followed by stimulus-evoked activity were repeated. We applied either 1) DNQX (0.5-1 μM) + AP5 (10-20 μM) or 2) Picrotoxin (PTX, 5 μM). The stimuli consisted of brief electric shocks delivered through a single electrode of the array located near layers 2/3 of the cortex culture. Each stimulus had a bipolar time course with amplitude –S for 50 μs in duration followed by an amplitude at +S/2 for 100 μs. We tested three similar sets of stimuli, each with 10 different stimulus levels, in μA (S1=10, 20, 30, 50, 60, 80, 100, 120, 150, 200; S2=6, 12, 24, 50, 65, 80, 100, 125, 150, 200; or S3=6, 12, 24, 50, 74, 100, 150, 200, 300, 400.) Different amplitudes were applied in pseudorandom order at 5 s intervals. MEA recordings in monkey. All procedures were in accordance with NIH guidelines and were approved by the NIMH Animal Care and Use Committee. 96-channel MEA (10x10 grid with no corner electrodes, 400 μm separation, and 1.0 mm electrode length; BlackRock Microsystem) were chronically implanted in the right arm representation region of pre-motor cortex of two monkeys (Macaca mulatta, adults, one male, one female). Ongoing activity was recorded for 30 min. Monkeys were awake, but not engaged by any task or controlled sensory stimulation. Extracellular signals were sampled at 30 kHz and filtered off-line (1 to 100 Hz; phase neutral, 4th order Butterworth). For Figure 5B, a 4x4 subset of electrodes was analyzed, matching spatial dimensions of the 4x4 coarse resolution view of the in vitro data (see below). For Figure 5C, 4x4 patterns based on a coarse-binned set of 8x8 electrodes spanning larger spatial area were analyzed. MEA recordings in rats. Urethane anesthetized rats aged 15 – 25 days were studied. As described previously (Gireesh and Plenz, 2008), an MEA was inserted into superficial layers of barrel cortex (Michigan probe 8x4 electrodes, 200 μm interelectrode distance). Ongoing activity was recorded for 20 – 30 min (4 kHz sampling rate, 1 – 3,000 Hz bandwidth) and filtered between 1 – 100 Hz (phase neutral, 4th order Butterworth) to obtain the LFP. In Figure 5B, the 8x4 MEA was coarse binned into a 4x4 pattern with 2 electrodes contributing to each bit. In Figure 5C, 4x4 patterns based on only 16 electrodes were also analyzed. Population event detection. For each electrode, the standard deviation (SD) of the LFP was calculated. Automated (Matlab) population event detection entailed first identifying large (< -4SD in vitro, -3SD in vivo) negative LFP fluctuations. Second, the time stamp ti, amplitude ai, and electrode ei of each negative peak (nLFP) occurring during these large fluctuations was extracted. Next, consecutive nLFPs were assigned to the same population event if they occurred sufficiently close (ti+1 – ti < τ) in time. The time threshold τ was determined based on the inter-nLFP-interval distribution, as described previously (Shew et al., 2009). 4 Event size distribution, neuronal avalanches, and κ. We defined the size s of a population event as the sum of all ai that occur during the event. During one recording we typically measured thousands of population events. We used a statistical measure called κ to quantify the character of each recording. The measure κ was developed to provide a network-level gauge of the E/I ratio, as demonstrated in a previous study (Shew et al., 2009). Similar to the Kolmogorov statistic, κ is computed based on the cumulative probability distribution of event sizes, F(β), which describes the probability of observing an event with size less than β. More specifically, κ quantifies the similarity of the measured F(β) and a theoretical reference distribution FNA(β). The reference distribution was chosen based on empirical observations that cortical networks with unaltered E/I tend to generate a probability distribution of population event sizes, P(s)~s-1.5, which defines neuronal avalanches (Beggs and Plenz, 2003; Shew et al., 2009). Importantly, population events are not considered to be neuronal avalanches unless they have this statistical property. Thus, FNA(β)~β-0.5 is the cumulative distribution corresponding to neuronal avalanches and κ assesses how close the observed distribution is to that of neuronal avalanches. By definition, κ is 1 plus the sum of 10 differences between F(β) and FNA(β). += k 1 1 10 (cid:229) 10 = i 1 NA F ( b ) i - F ( b ) i . (1) Therefore, κ≈1 indicates a good match with the reference distribution and the existence of neuronal avalanches. Positive (negative) deviations away from κ≈1 indicate a hyper- (hypo-) excitable network and the absence of neuronal avalanches. The 10 points at which the differences are computed are logarithmically spaced over the range of measured event sizes. It was previously shown that κ is not sensitive to changes in the number of differences (5 to 100 were tried; Shew et al., 2009). Binary patterns and entropy H. Each population event is represented by an 8x8 binary pattern with one bit per recording electrode. A bit is set to 1 if the corresponding electrode is active during the event; otherwise it is set to 0. The entropy of this set of patterns is defined as -= H (cid:229) n = i 1 p i log 2 p i , (2) where n is the number of unique binary patterns and pi is the probability that pattern i occurs. In Figure 2B, the black curve is based on the 8x8 binary patterns which represent the 60 electrodes of the MEA. We also studied entropy at different spatial resolutions by coarse-binning and for different spatial extents by using sub-regions of the MEA. The green curves in Figs. 2D, 3B, and 4A were obtained by reducing spatial resolution through coarse-binning of 8x8 patterns into square 4x4 patterns. Each bit in the 4x4 pattern was dependent on the state of 4 neighboring 2x2 electrode sets; if at least one electrode was active the bit was set to 1. Reduced spatial extent was tested with16-bit patterns based on only 16 electrodes arranged in a 4x4 square (Fig. 2D). As pointed out in the results, these 4x4 patterns also reduce potential undersampling bias when compared to 8x8 patterns. For stimulus-evoked activity, binary patterns were defined based on LFP activity 5 measured during 500 ms following the stimulus. If the measured response at an electrode exceeded -8SD of the noise, then the corresponding bit was set to 1, otherwise it was set to 0. The stimulation electrode was always set to 0. Note that the lack of corner electrodes on the MEA means that the corner bits of 8x8 patterns are always zero. This implies that the maximum entropy we could possibly record for 8x8 patterns is 260 rather than 264. For coarse-binned 4x4 patterns, the likelihood that the corner bits are active is slightly lower (about 25% lower). These effects are present for all E/I ratios examined. Therefore, they may affect the absolute values of entropy measurements, but they are not important for our conclusions, which are primarily based on changes in entropy. This is further confirmed by the robustness of our results to selecting 4x4 subregions from the center of the MEA for which corner electrodes are not missing (Fig. 2D). The calculation of entropy entails estimating the occurrence probability for each pattern. Therefore, H generally will depend on number N of observed patterns unless N is so large that the probability of each pattern is well represented by the samples recorded. H will be underestimated for sufficiently small N but becomes independent of N for sufficiently large N. To estimate potential ‘undersampling bias’ we computed corrected values following the quadratic extrapolation method (e.g. Magri et al., 2009). First, we randomly selected a fraction f of samples from the full set of N patterns. We recomputed the entropy for fractions f=0.1 to 1 in steps of 0.1. We repeated this 10 times for each f. Next, we fit the average H versus f data with the following function: fH ( ) = H 0 - a fN - b fN ( ) 2 (3) The fit parameter H0 is the estimated corrected value reported in the results section. Mutual information MI. From a set of N binary patterns, we defined a participation vector qi (length N) for each recording site i. qi(j) =1 or 0 indicated that site i was active or inactive during event j. The interaction between site i and site j was quantified by the mutual information (Rieke et al., 1997; Dayan and Abbott, 2001) of qi and qj defined as qqMI ; ( i j ) = (cid:229) (cid:229) a b 1,0 1,0 qp ( i = qa , j = b log) 2 (cid:230) (cid:231) (cid:231) Ł qp ( qp ( i = qa , j i = qpa ( ) = b = j ) b ) (cid:246) (cid:247) (cid:247) ł , (4) where p(x) is the probability of x, p(x,y) is the joint probability x and y. MI quantifies (in bits) the information shared by the two sites and provides similar information as a cross correlation (Supplementary Fig. S1). The MI values reported in Results were averages over all pairs of sites, MI = 2 MM ( (cid:229)(cid:229)- M M 1 i qqMI ( ; = += i j i 1 1 ) . j - )1 (5) For the in vitro experiments, we also used mutual information in a different way to 6 quantify the efficacy of information transmission between stimulus and response. Here we computed MI(S;R) = H(R) - H(RS). H(R) is the entropy of the full set of response patterns for all stimuli. H(RS) is the conditional entropy, i.e. the response entropy for single stimuli, averaged across the different stimuli (Rieke et al., 1997; Dayan and Abbott, 2001). Likelihood of participation L Likelihood of participation Li for site i was defined as the fraction of patterns in which the site participated: = L i 1 N (cid:229) N = j 1 jq )( i . (6) The average likelihood of participation L for all M sites is discussed in the main text and defined as = L 1 NM (cid:229)(cid:229) M N = = i j 1 1 jq )( i . (7) Data shuffling to destroy interactions. For the purpose of understanding how the entropy changes due to interactions between sites, we created surrogate data sets by shuffling the events in which sites participated. The 1s and 0s were randomly reordered in each participation vector qi such that interactions between sites were destroyed, but Li and N remained fixed. Model. The model consisted of M=16 binary sites (1=active, 0=inactive). Each site was intended to model the activity of a large group of neurons like the nLFP recorded at an electrode in the experiments. The strength of interactions between site i and site j was modeled as an activation probability; pij was the probability that site i would be activated at time t+1 if site j was active at time t. Therefore, if a set J(t) of sites were active at time (cid:213) -= - p p 1 1( ) t then site i would be active with probability . Increasing iJ ij t )( tJj )( (decreasing) the E/I ratio was modeled by increasing (decreasing) the average pij in the range between 0.1/M and 1.5/M (all pij were initially drawn from a uniform distribution on [0,1] and then all were scaled down by dividing by a constant). Population events were modeled by activating a single initial site (like an electrical shock applied at a single electrode or a spontaneous activation in the experiments) and recording the resulting activations that propagated through the network. These dynamics were defined by ],)( [ ] [ (cid:213) -Q= Q=+ t )1 1 tJj )( where si(t) was the state of site i at time t, Θ[x]=0 (1) for x<0 (>0), and ζi(t) was a random number drawn from a uniform distribution on [0,1] at each update of each site. The states of all sites were updated simultaneously at each time step. Each population event in the model was represented with a 16 bit binary pattern (1= the site was active at least once during the response to the stimulus, otherwise 0). By simulating 1000 population events (always initiated at the same site), we generated a set of patterns for which the entropy was computed. From the event size distribution of network events we ts ( i z i z i t )( p ij ) - 1( - p iJ t )( - (8) 7 computed κ (see Shew et al., 2009). The event size was defined as the sum of all activations from all sites during the population event. The range of average pij values studied with the model resulted in a range of κ=0.6 to 1.6. Statistical Analysis. For determining the statistical significance of differences in entropy for different drug conditions and differences in κ for different drug conditions, we first used a one-way ANOVA to establish that at least one drug category was different from at least one other. Next we performed a post-hoc test of significant pairwise differences between the drug categories using a t-test with the Bonferroni correction for multiple comparisons. The same procedure was used to assess significance of differences in H and MI for different categories of κ. RESULTS In all of our experiments, multi-electrode array (MEA) recordings of the local field potential (LFP, Fig. 1A) were used to obtain patterns of cortical population activity. We defined a recording site as ‘active’ if it presented a large, negative deflection in the LFP (Fig. 1A, green). We have previously demonstrated that such negative LFP deflections correlate with increased firing rates of the local neuronal population for each of the experimental preparations studied here: superficial layers of organotypic cultures (Shew et al., 2009), urethane anesthetized rat (Gireesh and Plenz, 2008) and awake monkeys (Petermann et al., 2009). We define a ‘population event’ as a set of electrodes which were active together within a short time (Methods). In our analysis, each population event was represented by a binary spatial pattern with one bit per recording site and 1 or 0 indicating an active or inactive site respectively (Fig. 1B; top). For each one hour recording in vitro (n=47) or 30 min recording in vivo (n=4 monkey, n=6 rat), we typically observed 103 to 104 population events. First, we systematically explored a range of E/I conditions in cortex slice cultures. A reduced E/I ratio was obtained by bath-application of antagonists of AMPA- and NMDA- glutamate receptor mediated synaptic transmission (DNQX, 0.5-1 μM; AP5, 10-20 μM). This resulted in population events that were typically small in spatial extent (Figs. 1B and 1C, left). Conversely, an increased E/I ratio was obtained with an antagonist of fast GABAA-receptor mediated synaptic inhibition (Picrotoxin; PTX, 5 μM), which led to stereotyped, spatially extended population events (Figs. 1B and 1C, right). In contrast, unperturbed E/I (Figs. 1B and 1C, middle) typically yielded a diverse pattern repertoire. The raster plots in Figure 1B (bottom) display examples of 100 consecutive population events recorded under the three different E/I conditions. Figure 1C displays example probability distributions of population event sizes for the three E/I conditions. We performed 11 recordings with reduced AP5/DNQX, 26 with no drug, and 9 with PTX. For each recording, we measured both ongoing activity and stimulus-evoked activity. For all recordings, we assessed the information capacity by computing the Shannon entropy of the full set of recorded binary patterns (Shannon, 1948; Rieke et al., 1997; Dayan and Abbott, 2001; see also Methods). 8 Figure 1: Measuring the neural activation pattern repertoire for a range of E/I conditions. A Example LFP recordings under conditions of reduced E (left), unperturbed E/I (middle), and suppressed reduced I (right). Scale bars: 250 ms x 10 μV (left, middle) and 250 ms x 100 μV (right). Population events were defined based on large negative deflections (<-4SD, green). B (top) Single examples of population events represented as binary patterns: 1=active site, 0=inactive. (bottom) Rasters including 100 consecutive population events represented as binary patterns; each row represents one event, each column represents one recording site. Left: reduced E. Middle: unperturbed. Right: reduced I. C Shape of event size distributions reveal changes in E/I, which are quantified with κ (see methods; broken line: power law with exponent of -1.5). Peak information capacity of ongoing activity for intermediate E/I and neuronal avalanches. Our first finding was that the entropy H for ongoing activity peaks at an intermediate E/I ratio. This was demonstrated with two different approaches. First, we compared entropy to the three pharmacological categories: AP5/DNQX, no drug, and PTX. We found that under the unperturbed E/I condition the average H was significantly higher than either the reduced E/I condition of the AP5/DNQX or the increased E/I condition of PTX (Fig. 2A, ANOVA, p<0.05). Second, we compared entropy to a previously developed statistical measure called κ, which characterizes E/I based on population dynamics of the 9 network (Shew et al., 2009; Methods). An advantage over the three pharmacology categories is that κ is a graded measure, thus providing a continuous function of entropy H versus E/I. A detailed definition of κ is given in Methods. Briefly, κ quantifies the shape of the population event size distribution, which is sensitive to changes in E/I (Fig. 1C): κ <1 indicated reduced E/I and κ >1 indicated increased E/I (Fig. 2B). Indeed, κ was significantly different for the two pharmacological manipulations compared to the no- drug condition (Fig. 2B; p <0.05). When we plotted entropy versus κ (Fig. 2C), we discovered a peaked function with maximum entropy occurring for κ≈1. This confirms our finding of peak entropy for the no drug condition (Fig. 2A) and provides a more refined view of the data; the peak occurred at κ*=1.16–0.12 (mean–SD, uncertainty determined by rebinning the experimental data, see Supplementary Fig. S2). The statistical significance of the peak in H was confirmed by comparing H for the ten experiments with κ closest to 1 with the ten experiments with smallest κ and ten with largest κ (p<0.05). In addition to providing a graded measure of E/I, κ assesses the statistical character of ongoing cortical population dynamics. Specifically, κ≈1 is the signature of neuronal avalanches (Shew et al., 2009), a type of population dynamics defined by a power-law population event size distribution with a power-law exponent near -1.5 (Beggs and Plenz, 2003; Stewart and Plenz, 2006; Gireesh and Plenz, 2008; Petermann et al., 2009; Shew et al., 2009). The computation of κ entails first computing the difference between a measured event size distribution and a theoretical reference distribution defined as a power-law with exponent -1.5 (Fig. 1C, green dashed). Next, this difference is added to 1 (for historical reasons, Shew et al., 2009), resulting in κ=1 for an exact match with a -1.5 power law, i.e. neuronal avalanches. In this context, our findings indicate that entropy is maximized under conditions which result in neuronal avalanches. Next we tested the robustness of the peak in H with respect to changes in spatial and temporal scales of recordings. First, as shown in Figure 2D (green), we found that the peak in H remained close to κ = 1 (κ*=1.01–0.02), even when the original 8x8 patterns were coarse-grained to obtain 4x4 patterns at half the spatial resolution (see Methods). Second, the peak was also maintained when the spatial extent of the recorded area was reduced by 75% (4x4 electrodes near center of the MEA; Fig. 2D). Finally, we confirmed that the peak persisted for a restricted recording duration of 12 minutes rather than one hour (Fig. 2D, purple). The robustness of our finding to shorter recording durations is important since estimations of entropy depend on the number of samples recorded (see below). 10 Figure 2: Ongoing activity - peak information capacity at intermediate E/I ratio specified by κ≈1. A Information capacity (entropy H) of the pattern repertoire is maximized for when no drugs perturb the E/I ratio. Significant differences marked with * (p<0.05). Box plot lines indicate lower quartile, median, upper quartile; whiskers indicate range of data, excluding outliers (+, >1.5 times the interquartile range). B The statistic κ, provides a graded measure of E/I condition based on network dynamics (methods). C Entropy H peaks near κ≈1. Each point represents one recording of ongoing activity (n=47, 8x8 MEA, 1 hr, color indicates drug condition; red=PTX, blue=AP5/DNQX, black=no drug). Line is the binned average of points. D The peak in entropy H is robust to changes in spatial resolution (green, 4x4 coarse-binned, 1 hr), spatial extent (orange, 4x4 subregion, 1 hr) and duration (purple, 4x4 coarse- binned, 12 min) of recording. (black, same data as in A). Error bars indicate mean±s.e.m. Peak information transmission between stimulus and response for intermediate E/I and neuronal avalanches. We now present measurements of stimulus-evoked activation patterns. A priori, one can expect a different distribution of stimulus-evoked patterns compared to ongoing activity and thus different entropy. Indeed studies suggest that ongoing activity is more diverse than typical stimulus-evoked activity (Welicky, 2004; Luczak et al., 2009; Churchland, 2010). However, if the entropy of evoked patterns changes with E/I in the same way that we found for ongoing activity, then evoked entropy may also peak near κ=1. This possibility is in line with significant evidence that ongoing activity in the cortex is intimately related to stimulus-evoked activity (Kenet et al., 2003; Ji and Wilson, 2007; 11 Han et al., 2008; Luczak et al., 2009). For instance, stimulus-evoked activity patterns recur during ongoing activity, both at the population level (Kenet et al., 2003; Han et al., 2008) and the level of spike sequences (Ji and Wilson, 2007). Therefore, our next aim was to test whether our finding of peak entropy near κ=1 also holds for stimulus-evoked activity. Stimuli consisted of 10 different amplitude single bipolar shocks each applied 40 times in randomized order though a single electrode of the MEA within cortical layers II/III (Methods). A binary pattern was constructed to represent each response during the 20 – 500 ms after the stimulus. The evoked entropy H was calculated for the set of 400 stimulus-evoked activation patterns for each E/I. As found for ongoing activity, the evoked entropy was highest near κ ≈1 for both fine and coarse spatial resolution (Fig. 3A; black - 8x8, green - coarse-grained 4x4, p<0.05). In the introduction, we gave a simple example in which information transmission from input to output was limited due to low entropy. With our measurements of network responses (i.e. output) to stimuli (i.e. input), we can directly test whether efficacy of information transmission is optimized when entropy is maximized. This idea is concisely summarized in the following equation: MI(S;R) = H(R) - H(RS). Here, MI(S;R) is the mutual information of stimulus and response which quantifies the information transmission (Rieke et al., 1997; Dayan and Abbott, 2001). H(R) is the entropy of the full set of response patterns for all stimuli, while H(RS) is the conditional entropy, i.e. the average entropy per stimulus (Rieke et al., 1997; Dayan and Abbott, 2001). As shown above, H(R) is maximized near κ ≈1. Since, H(RS) is always positive, MI(S;R) is bounded by H(R), and thus potentially also peaks near κ =1. Indeed, we measured MI(S;R) under different E/I conditions and found that stimulus-response mutual information was maximized near κ ≈1 (Fig. 3B; black - 8x8, green - coarse-grained 4x4, p<0.05). Figure 3: Stimulus-evoked activity - peak information transmission at intermediate E/I ratio specified by κ≈1. A Single shock stimuli with 10 different amplitudes (10-200 μA) were applied 40 times each using a single electrode. The pattern repertoire of stimulus-evoked activity has maximum 12 entropy near κ≈1. This holds for 8x8 response patterns (black line) as well as coarse resolution 4x4 patterns (green line). Points correspond to 8x8 patterns: light blue – AP5/DNQX, gray – no drug, pink – PTX. B The efficacy of information transfer, i.e. mutual information of stimulus and response, also peaks near κ≈1. The dashed line indicates the highest possible mutual information given 10 stimulus levels. (black - 8x8; green – 4x4). Error bars indicate s.e.m. Competition between activity rates and interactions explains peak in entropy. To identify and quantify the mechanisms leading to the peak in entropy near k=1, we analyzed in more detail the coarse-grained 4x4 patterns measured during ongoing activity (Fig. 2D, green). A priori, the total number of unique patterns that are possible is 216, implying a maximum H≤log2(216)=16 bits. This maximum would be reached if all 216 patterns occurred with equal probability. However, during a 1 hr recording, the network did not generate all possible patterns, nor were different patterns equally likely, resulting in H that was always below 16 bits. The peak in H was explained by three main factors that changed with the E/I ratio: i) the number N of patterns observed during the recording, ii) the likelihood L that sites participate in patterns, and iii) the strength of interactions between sites. The first two effects are related to the rates of observed activity and impose upper bounds on H: effect i requires H ≤ log2(N) (dash-dot line in Figs. 4A) and effect ii limits H in a way that depends on L (dashed line Figs. 4A). Specifically, the highest possible entropy for a given L can be computed by assuming that sites are independent, H ( (cid:229) -< M = i 1 L i log 2 L i -+ 1( L i log) 2 - 1( L i ) ) , (9) where M is the number of recording sites and Li is the likelihood of participation for site i. This formula is based on the fact that the entropy of two independent systems combined is the sum of their individual entropies. Since a single site i is either active (with probability - -- - Li) or inactive (with probability 1-Li), its entropy is L L L L . log log) 1( ) 1( i i i i 2 2 Thus, adding the entropy of all sites, we obtain the formula above. When L<1/2, increasing L increases the upper bound on H. When L>1/2, increasing L decreases the upper bound on H. We found that L increased over the range of E/I conditions we studied (Figs. 4C), while the number of patterns N did not show a systematic trend. We turn now to effect iii. Increased interactions between sites always reduce H due to the increased redundancy of the information at different sites (Schneidman et al., 2003). We found that site-to-site interactions during ongoing activity increased with E/I (Figs. 4E), and quantified this trend in two ways. First we computed mutual information (MI) between the activity recorded from different pairs of sites (Figs. 4E; red). Note that above we used mutual information in a different way, computed between stimulus and response MI(R;S) to assess information transmission. Second, we estimated the effect of interactions by computing the drop in entropy resulting from shuffling the data. The shuffling procedure destroyed interactions by randomizing the set of population events in 13 which each site participated, while keeping L and N fixed (Methods). The entropy of the shuffled data for the corresponding original κ value is shown in figure 4A (black) and, as expected, nearly reached the bounds set by the combined effects i and ii. The difference in entropy DH between the measured and shuffled data is due to interactions (Figs. 4E, blue). DH has previously been used to quantify redundancy (Dayan and Abbott, 2001). In summary, at low E/I, effects ii and iii compete and effect ii wins, i.e. activity rates drop sufficiently low to cause low entropy even though interactions are also low. At high E/I, effects ii and iii cooperate, i.e. both high activity rates and strong interactions cause low entropy. Entropy peaked at an intermediate E/I ratio at which interactions between sites were not too low or too high (specified by MI≈0.2) and activity was not too depressed (L≈0.25). Figure 4: Peak information capacity explained. A detailed analysis of in vitro experimental results (left, Fig. 2b, green) and model results (right, Fig. 2c, blue) A,B Upper bounds on entropy are set by 1) the average likelihood that sites participate in patterns (dashed) and 2) the number of patterns observed (dash- dot). When the effects of interactions are removed by shuffling (methods), the entropy reaches these bounds (black), but the measured entropy (green) is always lower due to interactions. C,D Rise in participation likelihood L as E/I ratio is increased. This rise accounts for the bounds (dashed) shown in A,B. E,F Rise in interactions between sites (mutual information, red) is proportional to the loss in information capacity ΔH (blue). All error bars indicate s.e.m. We remark that, if N were large enough (e.g. for longer recording duration), the upper bound due to effect i would become irrelevant, in which case, we still expect H to peak near κ=1 due to the combined effects of interactions (ΔH) and L. Nonetheless, the persistence of the peak in H for shorter duration recordings may be more relevant for 14 cortex operations which occur on shorter time scales. We also tested the extent to which our measurements are impacted by sample size following the methods developed by Magri et al. (2009). The difference between our measured H and ‘corrected’ H was 0.06–0.06 (mean±SD) bits for 4x4 in vitro ongoing activity patterns and 0.22–0.18 bits for the 8x8 patterns. Thus, sample size effects are small compared to the variability from one experiment to another (see error bars in Fig. 2.) We also point out that N, L, and MI are not the only factors that could potentially influence H. For example, not every site was equally likely to be active. Such spatial structure is expected to decrease entropy compared to a spatially homogeneous system with all other properties held fixed. This was not a major influence in our results. Experimental results confirmed in a computational network-level model. To gain further insight on our experiments, we compared our results to a network-level simulation, which has been used previously to model neuronal avalanches (Haldeman and Beggs, 2005; Kinouchi and Copelli, 2006; Shew et al., 2009). The model consisted of 16 binary sites. The state (1=active, 0=inactive) of each site was intended to represent a population of neurons in the vicinity of a recording electrode (Methods). The propagation of activity from one site to another was treated probabilistically; a connection matrix p with entries pij specified the probability that site i would become activated due to site j having been activated in the previous time step. Increases (decreases) in E/I were modeled by increasing (decreasing) the average pij value through the range 0.006 to 0.1. For each ‘E/I condition’, 1000 population events were simulated, beginning with a single initially active site and the resulting patterns of activity were recorded. To facilitate comparison with our experimental results we also parameterized each E/I condition of the model using k, based on population event size distributions. In good agreement with our experiments, we found that entropy reached a peak for k≈1 (Fig. 4B; green). Moreover, the explanation of peak entropy in terms of the competition between activity rates and site-to- site interactions also held for the model. Just as in the experiments, when the model data was shuffled to remove effects due to interactions, H (Fig. 4B; black) approached the upper bounds set by the number of events (Fig. 4B; dash- dot) and the likelihoods of participation (Fig. 4B; dashed). The model H results matched the experimental values, because the underlying changes in L versus κ (Fig. 4D) and the changes in site-to-site MI versus κ (Fig. 4F, red) were very similar to those measured experimentally. This agreement is not trivial; the same values of entropy could in principle be reached with different combinations of the underlying L and MI versus κ. For example, a peak in H could result if L remained fixed at 0.5 and interactions were minimized at κ=1. Site-to-site mutual information in the model reached slightly lower levels for high κ when compared to experiments(Fig. 4F, red), which could be due to the lack of significant structure in the model connectivity matrix p. 15 In vivo entropy matches in vitro prediction. Finally, we analyzed recordings of ongoing activity from superficial cortical layers in two awake monkeys (premotor cortex) not engaged in any particular task and in urethane-anesthetized rats (n=6, barrel cortex) with no whisker stimulation. In agreement with previous studies (Gireesh and Plenz, 2008; Petermann et al., 2009), we found that the ongoing activity was organized as neuronal avalanches (Fig. 5A). More precisely, we found that κ=1.02–0.02 for the monkeys and κ=1.08–0.02 for the rats. Based on our in vitro findings, these κ values suggest that the in vivo networks are operating under E/I conditions that maximize entropy and information transmission. Although we cannot fully test this idea without a full range of κ in vivo, we can test whether the in vivo values of H, L, and MI match with those predicted from the in vitro results. As shown in Figure 5B and summarized in Table 1, we found good agreement with these predictions. We found no statistically significant difference between the in vivo results and the prediction from in vitro experiments with the same range of κ (1.0<κ<1.1, p<0.05). Nonetheless, the fact that entropy values in vivo were slightly higher than the in vitro results, may be due to the corresponding slightly lower MI values. Figure 5: In vivo properties predicted from in vitro results. A Population event size distributions from ongoing activity in two awake monkeys (blue) and an example rat (green) are near a power-law with exponent -1.5 (dashed line), i.e. they exhibit neuronal avalanches and κ≈1. B In line with in vitro and model predictions for κ≈1, in vivo entropy 16 was high and mutual information between recording sites was moderate. (stars - two recordings on different days from each monkey; squares - anesthetized rats, n=6). The spatial extent of recorded area was approximately matched. C The result holds even when the spatial scales and resolution differ by factor of 4. We note that the in vivo values of MI, which are based on LFP measurements, coexist with low values of pair-wise correlation r between spiking activity of units (mean±s.e.m. r=0.03±0.01, Supplementary Fig. S3), in line with recent reports for awake monkeys (Ecker et al., 2010) and anesthetized rats (Renart et al., 2010). The success of our prediction requires matching the number of recording sites (16 here), but is robust to large changes in spatial extent and resolution of recordings (Fig 5C). The prediction is also robust to changes in the threshold used for generating binary activity patterns from continuous LFP data (Supplementary Table S1). Entropy, H (bits) Participation likelihood, L Site-to-site mutual information, MI (bits) In vitro predictions for 1.0<κ<1.1 Awake monkeys κ=1.02–0.02 Anesthetized rats κ=1.08–0.02 5.7±1.6 7.5–0.5 7.1–1.2 0.3±0.1 0.3–0.03 0.4–0.1 0.2±0.2 0.1–0.01 0.2–0.1 Table 1: In vivo results match in vitro predictions. Given the range of κ found in the in vivo recordings (1<κ<1.1), our in vitro results provide the predictions of H, L, and MI shown in the first row. The corresponding measurements from the awake monkeys (second row) and anesthetized rats (third row) match the in vitro predictions, i.e. they are not significantly different (p<0.05). Corresponding data are shown in Fig. 5B. All numbers are mean±SD. DISCUSSION We employed in vitro and in vivo experiments as well as a computational model to study the effects of the E/I ratio on entropy and information transmission in cortical networks. We analyzed multisite measurements of LFP recorded during ongoing as well as stimulus- evoked activity. We found that entropy and information transmission are maximized for the particular E/I ratio specified by κ=1, which is the same E/I condition under which neuronal avalanches emerge. We emphasize that the relative changes in H as we altered E/I are the meaningful results of our in vitro study; the absolute entropy values in bits depend upon arbitrary aspects of the analysis and measurements, e.g. the number of electrodes in the MEA. Thus, we are not suggesting that there is an absolute cap on the information that a cortical circuit can 17 represent at ~10 bits and it is not appropriate to compare our H values to those found in other studies of population entropy measures (e.g. Quian Quiroga and Panzeri, 2009). The important feature of our result is the peak in H near κ ≈1. We expect that any measure of population entropy would also peak for the same intermediate E/I, specified by κ ≈1. Previous studies have separately addressed the topics of entropy maximization (Laughlin, 1981; Dong and Atick, 1995; Dan et al., 1996; Li, 1996; Rieke et al., 1997; Dayan and Abbott, 2001; Garrigan et al., 2010), neuronal avalanches (Beggs and Plenz, 2003; Haldeman and Beggs, 2005; Stewart and Plenz, 2006; Ramo et al., 2007; Gireesh and Plenz, 2008; Tanaka et al., 2009; Petermann et al., 2009; Shew et al., 2009), and the balance of E/I (van Vreeswijk and Sompolinsky, 1996; Shadlen and Newsome, 1998; Shu et al., 2003; Okun and Lampl, 2008; Susillo and Abbott, 2009; Roudi and Latham, 2007), but our work is the first to show how these ideas converge in cortical dynamics. Significant evidence suggests that maximization of entropy is an organizing principle of neural information processing systems. For example, single neurons in the blowfly visual system have been shown to exhibit spike trains with maximized entropy, considering the stimuli the fly encounters naturally (Laughlin, 1981). Applied at the level of neural populations, the principle of maximized entropy has provided successful predictions of receptive field properties in mammalian retina (Garrigan et al., 2010), lateral geniculate nucleus (Dong and Atick, 1995; Dan et al., 1996), and visual cortex (Li, 1996). Our work shows that the potential ability of a neural population in the cortex to achieve maximum entropy and maximum information transmission depends on the E/I ratio. Thus, if such properties are optimal for the organism, then the particular E/I ratio specified by κ=1 may best facilitate this goal. We note that our investigation is not directly related to ‘maximum entropy’ models (e.g. Schneidman et al., 2006). In those studies, the aim was to use the maximum entropy principle (Jaynes, 1957) to find the simplest model to describe an experimental data set; entropy served as a modeling constraint. In contrast, here we compare the entropy across different experiments, searching for conditions which result in maximum entropy; entropy measurements are the results. Several theory and modeling studies (including our own model) offer a deeper explanation of why κ=1 and neuronal avalanches occurs under E/I conditions which maximize entropy and information transmission (Beggs and Plenz, 2003; Haldeman and Beggs, 2005; Ramo et al., 2007; Tanaka et al., 2009). Recall that neuronal avalanches and κ=1, by definition, indicate a power-law event size distribution with exponent -3/2. This same property is found in many dynamical systems that operate near ‘criticality’. Criticality refers to a particular mode of operation balanced at the boundary between order and disorder (e.g. Stanley, 1971; Jensen, 1998), akin to the balance of excitation and inhibition that we explore in our experiments. In our model, criticality occurs when the average pij equals 1/M (M is the number of sites). When pc>1/M, activity propagation is widespread and highly synchronous, like a seizure, while pc<1/M results in weakly interacting, mostly independent neurons (Beggs and Plenz, 2003; Haldeman and Beggs, 2005; Kinouchi and Copelli, 2006). The balanced propagation that occurs at criticality might be attributed to 18 interactions between excitatory and inhibitory neurons in the cortex. Using theory of Boolean networks, Ramo et al. (2007) showed theoretically that entropy of the event size distribution is maximized at criticality. Simulations of a model similar to our own showed that the number of activation patterns that repeat is maximized at criticality (Haldeman and Beggs, 2005). Tanaka et al. (2009) found that recurrent network models in which information transmission is optimized also exhibit neuronal avalanches and repeating activation patterns. Likewise, it has been shown that mutual information of input and output in feed-forward network models is maximized near criticality (Beggs and Plenz, 2003) and mutual information between neurons in the same network is maximized at criticality (Greenfield and Lecar, 2001). In line with these theory and model predictions, our results are the first experimental demonstration of peak entropy and information transmission in relation to criticality in the cortex. Finally, a separate line of research has focused on the E/I ratio in cortical networks. Models emphasize the importance of balanced E/I for explaining the variability observed in spike trains (van Vreeswijk and Sompolinsky, 1996; Shadlen and Newsome, 1998), low correlations between spiking units (Renart et al., 2010), and generating diverse population activity patterns (Susillo and Abbott, 2009), which may play a role in memory (Roudi and Latham, 2007). Moreover, in vivo experiments have shown that synaptic input received by cortical neurons exhibits a fixed ratio of excitatory to inhibitory current amplitudes (Shu et al., 2003; Okun and Lampl, 2008). Since we measure κ≈1 in vivo, it follows that the ‘balanced E/I’ discussed in these previous studies may also correspond to the optimal E/I that we identify here. In summary, our results suggest that by operating at the E/I ratio specified by κ ≈1, the cortex maintains a moderate level of network-level activity and interactions which maximizes information capacity and transmission. This finding supports the hypotheses that balanced E/I and criticality optimize information processing in the cortex. REFERENCES Beggs JM, Plenz D (2003) Neuronal avalanches in neocortical circuits J Neurosci 23:11167-11177. Churchland MM et al. (2010) Stimulus onset quenches neural variability: a widespread cortical phenomenon. Nat Neurosci 13, 369-378. Dan Y, Atick JJ, Reid RC (1996) Efficient coding of natural scenes in the lateral geniculate nucleus: experimental test of a computational theory. J Neurosci 10:3351-3362. Dayan P, Abbott LF (2001) Theoretical neuroscience (MIT Press, Cambridge, Massachussetts, USA). Dichter M, Ayala G (1987) Cellular mechanisms of epilepsy: a status report. Science 237:157-164. Dong DW, Atick JJ (1995) Temporal decorrelation: a theory of lagged and nonlagged 19 responses in the lateral geniculate nucleus. Network: Computation in Neural Systems 6:159-178. Ecker AS et al. (2010) Decorrelated neuronal firing in cortical microcircuits. Science 327:584-587. Fiser J, Chiu C, Weliky M (2004) Small modulation of ongoing cortical dynamics by sensory input during natural vision. Nature 431, 573-578. Garrigan P et al. (2010) Design of a trichromatic cone array. PLoS Comput Biol 6:e1000677. Gireesh ED, Plenz D (2008) Neuronal avalanches organize as nested theta- and beta/gamma-oscillations during development of cortical layer 2/3. Proc Nat Acad Sci USA 105:7576-7581. Greenfield, E., and Lecar, H. (2001). Mutual information in a dilute, asymmetric neural network model. Physical Review E 63, 1-10. Haldeman C, Beggs JM (2005) Critical branching captures activity in living neural networks and maximizes the number of metastable states. Phys Rev Lett 94:058101. Han F, Caporale N, Dan Y (2008) Reverberation of recent visual experience in spontaneous cortical waves. Neuron 60:321-327. Ji D, Wilson MA (2007) Coordinated memory replay in the visual cortex and hippocampus during sleep. Nat Neurosci 10:100-107. Jacobs et al. (2009) Ruling out and ruling in neural codes. Proc. Nat. Acad. Sci. USA 106:5936-5941. Jaynes ET (1957) Information theory and statistical mechanics. Phys Rev 106:62–79. Jensen HJ (1998) Self-organized criticality: emergent complex behavior in physical and biological systems. (Cambridge University Press, Cambridge, UK). Kenet T, Bibitchkov D, Tsodyks M, Grinvald A, Arieli A (2003) Spontaneously emerging cortical representations of visual attributes. Nature 425:954–956. Laughlin S (1981) A simple coding procedure enhances a neuron’s information capacity. Zeitschrift fur Naturforschung 36:910-912. Li Z (1996) A theory of the visual motion coding in the primary visual cortex. Neural Comput 8:705-730. Luczak A, Barthó P, Harris KD (2009) Spontaneous events outline the realm of possible sensory responses in neocortical populations. Neuron 62:413-425. Magri C, Whittingstall K, Singh V, Logothetis NK, Panzeri S (2009) A toolbox for the fast information analysis of multiple-site LFP, EEG and spike train recordings. BMC 20 Neurosci. 10:81. Nauhaus I, Busse L, Carandini M, Ringach DL (2009) Stimulus contrast modulates functional connectivity in visual cortex. Nat Neurosci 12:70-76. Okun M, Lampl I (2008) Instantaneous correlation of excitation and inhibition during ongoing and sensory-evoked activities. Nat Neurosci 11:535-537. Petermann T et al. (2009) Spontaneous cortical activity in awake monkeys composed of neuronal avalanches. Proc Nat Acad Sci USA 106:15921-15926. Pola G, Thiele A, Hoffmann K-P, Panzeri S (2003) An exact method to quantify the information transmitted by different mechanisms of correlational coding. Network: Comput. Neur. Syst. 14:35-60. Quian Quiroga R, Panzeri S (2009) Extracting information from neuronal populations: information theory and decoding approaches. Nat Rev Neurosci 10:173-185. Ramo P, Kauffman S, Kesselia J, Yli-Harja O (2007) Measures for information propagation in Boolean networks. Physica D 227:100-104. Renart A et al. (2010) The asynchronous state in cortical circuits. Science 327:587-590. Rieke F, Warland D, de Ruyter van Stevenick R, & Bialek W (1997) Spikes (MIT Press, Cambridge, Massachussetts, USA). Roudi Y, Latham PE (2007) A balanced memory network. PLoS Comp Biol 3:1679-1700. Schneidman E, Berry II MJ, Segev R, Bialek W (2006) Weak pairwise correlations imply strongly correlated network states in a neural population. Nature 440:1007-1012. Schneidman E, Bialek W, Berry II MJ (2003) Synergy, redundancy, and independence in population codes. J Neurosci 23:11539 –11553. Shadlen MN, Newsome WT (1998) The variable discharge of cortical neurons: implications for connectivity, computation, and information coding. J Neurosci 18:3870- 96. Shannon CE (1948) A mathematical theory of communication. Bell System Technical J 27:379-423, 623–656. Shew WL, Yang H, Petermann T, Roy R, Plenz D (2009) Neuronal avalanches imply maximum dynamic range in cortical networks at criticality. J Neurosci 29:15595-15600. Shu Y, Hasenstaub A, McCormick DA (2003) Turning on and off recurrent balanced cortical activity. Nature 423:288-293. Stanley HE (1971) Introduction to Phase Transitions and Critical Phenomena (Oxford 21 University Press, New York, USA). Stewart C, Plenz D (2006) Inverted-U profile of dopamine-NMDA-mediated spontaneous avalanche recurrence in superficial layers of rat prefrontal cortex. J Neurosci 26:8148- 8159. Sussillo D, Abbott LF (2009) Generating coherent patterns of activity from chaotic neural networks. Neuron 27:544-557. Tanaka T, Kaneko T, Aoyagi T (2009) Recurrent infomax generates cell assemblies, neuronal avalanches, and simple cell-like selectivity. Neural Comput 21:1038-1067. van Vreeswijk C, Sompolinsky H (1996) Chaos in neuronal networks with balanced excitatory and inhibitory activity. Science 274:1724-1726. Supplementary Material for 'Information capacity and transmission are maximized in balanced cortical networks with neuronal avalanches' by Shew et al. Relationship between mutual information and correlation coefficient To quantify interactions between sites we used average pair-wise mutual information (MI). A more traditional approach is to use average pair-wise correlation coefficients (CC). Our reason for working with mutual information is two-fold. First, MI is less sensitive to noise when interactions are very weak. Second, MI arises from information theory and, thus, is a more natural fit with the study of entropy. Nonetheless, MI and CC are closely related. In Fig S1 we compare both quantities for the in vitro data presented in the main text. Also marked in Fig S1 are estimated theoretical bounds on the relationship between MI and CC. The lower bound is reached for L=0.5, while the upper bound corresponds to the extreme values of L=0 and L=1. 22 Figure S1 Relationship between mutual information (MI) and correlation coefficient (CC). The black points compare MI and CC for all in vitro experiments. The dashed line is an estimated lower bound, which is reached for L=0.5. The upper line is an estimated upper bound, which is reached for very low or very high L. The estimated bounds were obtained numerically for a single pair of binary vectors (10,000 events), each with the same L (0.1 to 0.9) and CC (0.1 to 0.9). These bounds are only approximate, because in the experiments, L is not the same from one electrode to another. Binning sensitivity for H vs. k. Here we show that the results shown in Fig 2C are robust to different choices of the bins used to produce the average line. The variability of the peaks of these curves was used to estimate the uncertainty in the conclusion that peak entropy occurs at k≈1. Figure S2 Results robust to bin choices. The results shown in Fig 2C (black, green) for H vs. k for 8x8 patterns (top) and coarse-binned 4x4 patterns (bottom) were recomputed with different averaging bins. The different bin partitions are shown below the curves with 23 corresponding colors. Pair-wise spike train cross correlation in the monkey Here we report the spike count cross correlation values between unit activity recorded simultaneously with the monkey LFP recordings discussed in the main text. The average correlation between unit signals is significantly lower than that between the population signals provided by the LFP. As shown in Fig S2, the mean, s.e.m., and distributions of correlation coefficients were in good agreement with recent reports from awake monkeys (Ecker et al., 2010) and anesthetized rats (Renart et al., 2010). Spike sorting was performed with Plexon offline spike sorter (V2.8.8). 66 and 40 well isolated units were found for monkey 1 and monkey 2 respectively. Three principal components (PCA), peak-trough amplitude, and nonlinear energy were used as the sorting features. We defined ‘well isolated’ as follows: in a 2-D projection of at least 2 of the sorting features the unit must have a mean which is strongly different from the mean of noise waveforms (p≤0.001, multivariate ANOVA). If more than one unit was recorded from the same electrode, the difference between means of each unit was also required to be significant at this strict level (p≤0.001, multivariate ANOVA). To compute spike count cross correlations between each pair of units recorded during ongoing activity we followed established methods (Renart et al., 2010). First, to obtain spike count vectors, the spike time stamps of each unit were 1) binned with 1 ms temporal resolution, 2) convolved with a Gaussian window with 50 ms width. The cross correlation coefficient was computed between all pairs (2145 pairs for monkey 1, 780 pairs for monkey 2) of spike count vectors. Figure S3 Histograms of pairwise correlation coefficients of unit activity in the awake monkeys. As reported previously (Ecker et al., 2010; Renart et al., 2010) the average across all pairs is near zero and positive. The histogram of all pairwise CC values for monkey 1 (left) and monkey 2 (right) are shown. The mean±sem CC values were 0.050±0.002 and 0.015±0.001 for monkey 1 and monkey 2 respectively. Robustness to event detection threshold In line with previous studies, our in vivo monkey results were robust to changes in the detection threshold for nLFPs. For thresholds -2.5, -3 and -3.5 SD we found no significant changes as shown in Table S1 below (mean±s.e.m.). 24 Monkey 1 Day 1 Day 2 Monkey 2 Day1 Day 2 κ H MI 1.02±0.02 1.07±0.03 7.96±0.35 7.58±0.32 0.11±0.01 0.14±0.02 0.99±0.02 1.00±0.02 6.99±0.10 7.51±0.08 0.13±0.01 0.11±0.02
1909.05908
1
1909
2019-09-09T18:13:41
Neurological Nature of Vision and Thought and Mechanisms of Perception Experiences
[ "q-bio.NC" ]
Understanding of the phenomena of vision and thought require clarification of the general mechanism of perception. So far, philosophical inquiries and scientific investigations have not been able to address clearly the mysteries surrounding them. The present work is an attempt to unravel the essences of these phenomenal based on the presumption of computational brain. Within this context, the natures of thought is clarified, and the basis of the experience of perception is established. And by drawing from the successes of the developed tactile vision substitution systems (TVSS), which render some measure of vision,in vision handicapped persons, early or congenital blinds, the true nature of vision as cutaneous sensations is also divulged. The mechanism of perception involves sensing of the stimuli, and autonomous engagement of brain neuronal complexity resolution patterns; that is the brain implicit embedded computational instructions. Upon commencement of the triggers, brain computations, which aso involve engaging body's biophysical feedback system, are performed; and the results are outputted as motor signals that render the realization of perception. However, this requires deployment of a perception medium; an interface. Given the nature of efferent signals, there must be a (known) bio-mechanical system interface, other than the body muscle and skeletal system, which performs the needed function: Considering the fact that the vocal system performs such task for verbalization of brain's synthesis of language, the possibility of its further role in the experience of thought and vision, in the form of mostly quiet (inaudible) recital of the related motor signals, is suggested.
q-bio.NC
q-bio
Journal of Neurology & Stroke Neurological Nature of Vision and Thought and Mechanisms of Perception Experiences Short Communication Retired LBNL (UCB) Scientist, USA Volume 4 Issue 5 - 2016 *Corresponding author: Jahan N Schad, PhD., Retired LBNL (UCB) Scientist, 376 Tharp Drive, Moraga, Ca 94556, California, USA, Tel: 925-376-4126; Email: Received: March 28, 2016 Published: April 19, 2016 Abstract Understandings of the phenomena of vision and thought require clarification of the general mechanism of perception, -the experience prompted by the (brain) efferent signals -as well as the clarification of the natures of the related afferent signals, which drive the mechanism. So far, philosophical inquiries and scientific investigations have not been able to address clearly the mysteries surrounding them. The present work is an attempt to unravel the essences of these phenomena based on the presumption of computational functioning of the brain, a concept supported by scientific consensus. Within this context, the nature of the thought is clarified, and the basis of the experience of perception is established. And by drawing from the successes of the tactile vision substitution system (TVSS) [1], -- which renders a measure of vision perception in vision handicapped, early or congenital blinds -- the true nature of the vision, as cutaneous sensation, is also divulged. The mechanism of perception --what renders it and where it occurs --involves sensing of stimuli, and, or, the autonomous engagement of brain inherent neuronal complexity resolution patterns; the implicit embedded computational instruction (codes). Upon commencement of such triggers, --of which one may not be necessarily aware --brain computations, which also involve engaging body's biophysiological feedback system, are performed; and the results are outputted as motor (efferent) signals that render perception. However, embedded in the process of the development of the experience of perception, is the deployment of a perception medium; an interface. Given the nature of the efferent signals, there must be a (known) biomechanical system interface, --other than the irrelevant body muscle and skeletal systems --which performs the needed function: Considering the fact that the vocal system performs such task for the verbalization of brain's synthesis of language expressions, the possibility of its further role in the experiences of thought and vision, in the form of mostly quiet (inaudible) recital of related signals, is suggested. Keywords: Vision; Thought; Neuronal computation; Simulation; Utterance interface; Language; Biolinguistic Understanding of the phenomenon of Vision, which is to know "how and where we see what we see," beyond the knowledge of the biophysiological and optical aspects of the eyes, and the brain modalities where the trigger signals are processed, has remained a mystery. Neurosciences' knowledge of the central nervous system, and the brain neurocomputational concepts, suggest that brain neuronal code (computational patterns) processing of eye-extracted afferent data (somehow) renders vision perception. However, this still falls short of a complete and convincing addressing of the above question; leaving the vision concept vague, as it has always been. Further, and very detailed understanding of the anatomy and physiology of the vision and the related processes [2], are not likely to provide the answer. We resolve this ambiguity by bringing to light the nature of the vision afferent) signals, their processing in the brain, and where the (brain) efferent vision signals are experienced. Submit Manuscript http://medcraveonline.com Hypothesis Same difficulty for the though, as to "what it is and where it happens," has also always held true; beyond platitude. The philosophical addressing of mental processes and scientific understandings of the brain and its functions has not helped to resolve the puzzle either. A step toward the resolution of the thought ambiguity can be found in the biologic theory of linguistics, which according to Chomsky [3] entails the presence of neuronal language construct in the brain, and the proposition of two related interfaces: The first is "the thought system which provides a place for the interpreted Internal (I) language mechanism synthesis of the structured expression in the brain;" and the second is the "vocal system, activated by the motorsensory neurons," which renders language vocalizations; whether it is referential as in the calls of animals, or verbalized as in humans. However, despite this enlightening concept, the dilemma about the overall nature of the thought and its system still remains. J Neurol Stroke 2016, 4(5): 00152 2/3 perception is same as cutaneous perception, and that there should not be any differences in the natures of their afferent signals. system which includes a pulsating patch on the skin or tongue. To address these mysteries, we focus on: Taking note of the fact that such subjects never experience a) The nature of the brain information processing schemes vision, neither in waking hours nor in dreams [7], the experience and the triggers which drives them; and of such perceptions seemed inexplicable: On the face of it, the patch should only create cutaneous perceptions. True that from b) On the nature of the brain outputs and the need for the patch location, massive and simultaneous amount of data biological interfaces. pulses are sent to the brain, however, this should only lead to functioning of the The presumption of computational the development of some matter (object) perception, likely that brain is based on the vast body of computational sciences and of the patch itself. However, the repeated experimentally verified neurosciences findings, in tandem with experimental works in phenomenon can be explained in the context of brain neuronal the area of information processing of the neurons [4]; and the net computational procedure which can engage any available successes of the brain inspired scientific neural networks in neural circuitry (brain modality), including those of vision, for developing some measure of human-like intelligence. Extending the processing of large amount of afferent data, such as those the general workings principal of the artificial neural nets to the of cutaneous nature sensed from the TVSS patches. Given the brain neuronal computations [5], is a very plausible assumption: visual perception experiences of the blind experimental subjects, In the scientific neural networks the resolution of complex rendered by the brain post processed efferent signals, the problems calls for increasing units (layers) of calculation nodes, inference would be that the difference between the natures of and verifiable [6] brain neuronal plasticity allows for engagement the vision and cutaneous afferent signals must not be in kind but of various available neuronal modalities (possibly in hierarchical only in intensity, which dictates the vision details. And this claim manner). Obviously brain neuronal net with the estimated is evinced by the hazy visual-like perceptions of blind subjects availability of many trillions of biological microprocessors is an due to still insufficiency of the tactile afferent data. Based on such unfathomable complex, intelligent parallel process computation observations the disruptive discovery is that: the nature of vision engine that is evolutionary perfected and configured for sustenance of life. Implications are its lightening speed, complexity resolution potential by virtue of its constructs, and the engrained As in the case of thought, perception experiences require learned schemes (as neural patterns), for handling all relevant an interface, a display venue for the related brain (output) natural phenomena; including vision and thought that life entails. efferent signals-- generated post computational processing of The process of natural phenomena resolution in the brain the environmental stimuli from the vision and tactile sensing: A begins with the receipt of the sensed stimuli, and/or with the known human biological systems interface must be serving this autonomous deployment of brain inherent computational function: Considering the fact that the vocal system performs neuronal patterns (implicit complexity resolution codes); such a task for the verbalization of brain's synthesis of language which is followed by the onset of necessary computations while expressions efferent signals, the possibility of its versatility for engaging body's biophysiological feedback system. Clearly, these expression of the thought, the tactile and the vision perception operations, due to the ever presence of triggers, are perpetual; experiences in the form of quiet (inaudible) -- often unaware -- and the streaming outputs, as motor (efferent) signals, render recital of related information, is suggested. continuous perceptions of various phenomena. Among them are the experience of thought, triggered by internal or external environmental elements; and the experience of vision, mostly A measure of validation can be found in the very comprehensive by the external environmental triggers during waking hours; and detailed experimental work on mirror neuron activities of which one may or may not be aware. However, embedded in performed by Keysers et al. [8,9]: They examined a phenomenon the process of the experience of perception, is the availability called "tactile sympathy." In these studies, the areas of motor of a medium for it; an interface. Given the nature of the efferent neurons activated in a group of subjects watching a movie of signals, there must be a known biomechanical system interface, a second group being very lightly touched on the skin were other than body muscle and skeletal systems, which performs the significantly similar to those who were actually being touched. needed function. Also the combined fMRI and TMS evidence of such sympathy, The vocal system, mentioned earlier, is a proven candidate: which is shown in the work of Alaert et al. [10], provides This interface is responsible for the vocalization of the language, additional credence to our hypothesis. We believe these results and occasionally of thought; the latter true for almost all. And this provide strong experimental support for the concept of the tactile experience of switching from thought to talk, metaphorically a nature of vision presented in this theory. gearshift, discloses perhaps disruptively, the immense possibility For further validation of theory, we suggest experiments in which specific vocal motor neuron activities are monitored in different groups of subjects, some normally sighted and others As the to natures of vision efferent signals, we made a seemingly with congenital blindness. Subjects would be monitored during important and unexpected discovery by critically examining the speech, thinking, writing and visual (or, in the case of blind experimental results of the tactile vision substitution system subjects, vibro-tactile) engagements. Vocal vibrations during (TVSS), demonstrated in the initial work of Bach-y-Rita et al. [1], conscious thinking would also be recorded. The results of such Published in Nature: The work had established the development experimentation would definitively either prove or refute the of vision-like perception in blind subjects, when fitted with the theories put forward in this work. Citation: Schad JN (2016) Neurological Nature of Vision and Thought and Mechanisms of Perception Experiences. J Neurol Stroke 4(5): 00152. DOI: 10.15406/jnsk.2016.04.00152 Validation of presence of dual mode to the vocal system, which allows for expressions of audible and inaudible thoughts. Neurological Nature of Vision and Thought and Mechanisms of Perception ExperiencesCopyright:©2016 Schad 3/3 References 1. Bach-y-Rita P, Collins CC, Saunders FA, White B, Scadden L (1969) Vision substitution by tactile image projection. Nature 221(5184): 963-964. 2. Schwarz SH (1999) Visual Perception. McGraw Hill, New York, USA. 3. Chomsky N (2007) On Language. The New Press, New York, USA. 4. Kandel ER, Schwartz JH, Jessell TM (2010) Principles of Neurosciences. McGraw-Hill, New York, USA. 5. Schad NJ (2016) Brain Neurological Constructs: The Neuronal Computational Schemes for Resolution of Life's complexities. J Neurol Neurophysiol 7(1): 356. 6. Edelman GM (1987) Neural Darwinism. The Theory of Neuronal Group Selection. Basic Books, New York, USA. 7. Edison, Tommy (2013) Intangible Concepts to a Blind Person. Youtube Recording. 8. Keysers C, Wicker B, Gazzola V, Anton JL, Fogassi L, et al. (2004) A touching sight SII/PV activation during the observation and experience of touch. Neuron 42(2): 335-346. 9. Keysers C, Gazzola V (2009) Expanding the mirror: Vicarious activity for actions, emotions and sensations. Curr Opin Neurobiol 19(16): 666-671. 10. Alaerts K, Swinnen SP, Wenderoth N (2009) Is human primary motor cortex activated by muscular or direction-dependent features of observed movements? Cortex 45(10): 1145-1155. Conclusion Visual and cutaneous stimuli sensations are (computationally) processed in the brain similarly; which are evinced by the development of vision perception in blind subjects, congenial or otherwise, fitted with tactile visual substitution (TVSS) systems. It is the scarcity of normal tactile sense data in blinds, which limits their proper perceptions of the environment. In case of normal eyesight, retinal neurons figuratively extend themselves by virtue of receiving rays of Photons which are environmentally modulated for the physical reality of the object from which they are reflected. Putting it simply, in the experience of vision we are being touched by the external world, while in cutaneous experience we are physically touching them. Brain's computational operations are also constantly triggered by beings, exposure to other life phenomena, which are resolved, and streamed as efferent signals for perception. The experiences of perceptions in response to senses stimuli must be realized at a venue, a biomechanical interface which expresses (displays) the corresponding post process brain efferent signals: Vocal system, which serves language verbalization, is seemingly the only such device which could offer this possibility. And this leads to the presumption that vision, thought and tactile perceptions are but mostly inaudible utterances at the vocal machinery. Perhaps this discovery would be disturbing to the poetic thought; on the face of it; however, knowing that seeing has more to it than sweep of glance, the "Thought" would be more incensed of its romantic implications! Thanks are due to Professor Noam Chomsky for his pioneering biolinguistic work and for the lecture that has served as an overall inspiration for me. Further thanks go to Dr. Roya Noorishad, who instigated my interest in neurosciences. Acknowledgement Citation: Schad JN (2016) Neurological Nature of Vision and Thought and Mechanisms of Perception Experiences. J Neurol Stroke 4(5): 00152. DOI: 10.15406/jnsk.2016.04.00152 Neurological Nature of Vision and Thought and Mechanisms of Perception ExperiencesCopyright:©2016 Schad
1801.03880
1
1801
2018-01-11T17:18:42
Measuring the Complexity of Consciousness
[ "q-bio.NC", "physics.bio-ph" ]
The quest for a scientific description of consciousness has given rise to new theoretical and empirical paradigms for the investigation of phenomenological contents as well as clinical disorders of consciousness. An outstanding challenge in the field is to develop measures that uniquely quantify global brain states tied to consciousness. In particular, information-theoretic complexity measures such as integrated information have recently been proposed as measures of conscious awareness. This suggests a new framework to quantitatively classify states of consciousness. However, it has proven increasingly difficult to apply these complexity measures to realistic brain networks. In part, this is due to high computational costs incurred when implementing these measures on realistically large network dimensions. Nonetheless, complexity measures for quantifying states of consciousness are important for assisting clinical diagnosis and therapy. This article is meant to serve as a lookup table of measures of consciousness, with particular emphasis on clinical applicability of these measures. We consider both, principle-based complexity measures as well as empirical measures tested on patients. We address challenges facing these measures with regard to realistic brain networks, and where necessary, suggest possible resolutions.
q-bio.NC
q-bio
Measuring the Complexity of Consciousness Xerxes D. Arsiwalla1 , 2 , 3 and Paul Verschure1 , 2 , 3 , 4 1 Institute for Bioengineering of Catalonia, Barcelona, Spain. 2 Barcelona Institute for Science and Technology, Barcelona, Spain. 3 Universitat Pompeu Fabra, Barcelona, Spain. 4 Instituci´o Catalana de Recerca i Estudis Avan¸cats (ICREA), Barcelona, Spain. {[email protected]} Abstract. The quest for a scientific description of consciousness has given rise to new theoretical and empirical paradigms for the investiga- tion of phenomenological contents as well as clinical disorders of con- sciousness. An outstanding challenge in the field is to develop measures that uniquely quantify global brain states tied to consciousness. In partic- ular, information-theoretic complexity measures such as integrated infor- mation have recently been proposed as measures of conscious awareness. This suggests a new framework to quantitatively classify states of con- sciousness. However, it has proven increasingly difficult to apply these complexity measures to realistic brain networks. In part, this is due to high computational costs incurred when implementing these measures on realistically large network dimensions. Nonetheless, complexity measures for quantifying states of consciousness are important for assisting clinical diagnosis and therapy. This article is meant to serve as a lookup table of measures of consciousness, with particular emphasis on clinical appli- cability of these measures. We consider both, principle-based complexity measures as well as empirical measures tested on patients. We address challenges facing these measures with regard to realistic brain networks, and where necessary, suggest possible resolutions. Keywords: Consciousness in the Clinic, Computational Neuroscience, Complexity Measures. 1 Introduction In patients with disorders of consciousness, such as coma, locked-in syndrome or vegetative state, levels of consciousness are assessed in the clinic through a battery of behavioral tests and neurophysiological recordings. In particular, these methods are used to assess levels of wakefulness (arousal) and awareness in patients [28], [27]. Such assessments have led to a two dimensional operational definition of consciousness for clinical purposes. Assessments of awareness use behavioral and neurophysiological (fMRI or EEG) protocols in order to gauge how patients perform on various cognitive functions. Assessments of wakefulness are based on metabolic markers (if reporting is not possible) such as glucose 8 1 0 2 n a J 1 1 ] . C N o i b - q [ 1 v 0 8 8 3 0 . 1 0 8 1 : v i X r a uptake in the brain, captured using PET scans [17]. As such a clinically-oriented definition of consciousness enables classification of closely associated states and disorders of consciousness into clusters on a bivariate scale with awareness and wakefulness on orthogonal axes. Under healthy conditions, these two levels are almost linearly correlated, as in conscious wakefulness (high arousal and high awareness) or in deep sleep (low arousal and low awareness). However, in patho- logical states, wakefulness without awareness can be observed in the vegetative state [28], while transiently reduced awareness is observed following seizures [16]. Patients in the minimally conscious state show intermittent and limited non- reflexive and purposeful behavior [20], [19], whereas patients with hemi-spatial neglect display reduced awareness of stimuli contralateral to the side where brain damage has occurred [30]. Given the aforementioned scales for labeling states and disorders of consciousness, the crucial question is how should one quantify awareness and wakefulness from neurophysiological data? This is particularly useful for non-communicative patients such as those in coma or states of min- imal wakefulness. For this reason, several dynamical complexity measures have been developed. In this article, we first describe theoretically-grounded com- plexity measures and the challenges one faces when applying these measures to realistic brain data. We then outline alternative empirical approaches to classify states and disorders of consciousness. We end with a discussion on how these two approaches might inform each other. 2 Measures of Integrated Information Dynamical complexity measures are designed to capture both, network topol- ogy as well as causal dynamics. The most prominent among these is integrated information, denoted as Φ. This was first introduced in [37] and is defined as the quantity of information generated by a network as a whole, over and above that of its parts, taking into account the system's causal dynamical interactions. This reflects the intuition going back to William James that conscious states are integrated, yet diverse. Φ seeks to operationalize this intuition in terms of complexity, stating that complexity arises from simultaneous integration and dif- ferentiation of the network's structure as well as dynamics. Differentiation refers to functional specialization of neural populations, while integration, as a com- plementary design principle, results in distributed coordination among neural populations. This interplay generates integrated yet diversified information be- lieved to support cognitive and behavioral states. The earliest proposals defining integrated information were made in [37], [36] and [34]. Since then, considerable progress has been made towards the development of a normative theory as well as applications of integrated information [12], [14], [35], [1], [29], [7], [8], [26], [33], [5]. The core idea of integrated information as a whole versus parts quantity has been formalized in several distinct information measures such as neural com- plexity [37], causal density [32], Φ from integrated information theory: IIT 1.0, 2.0 & 3.0 [34], [12], [29], stochastic interaction [39], [11], stochastic integrated information [14], [1], [9] and synergistic Φ [23], [22]. Table 1 summarizes these measures along with corresponding information metrics upon which they have been based. Table 1. Theoretical complexity measures alongside their corresponding information metrics. Integrated Information Measures Information Metrics Neural Complexity Causal Density Mutual Information (MI) Granger Causality (GC) Stochastic Interaction Kullback-Leibler Divergence (KLD) IIT 1.0 & 2.0 Stochastic Integrated Information KLD MI or KLD IIT 3.0 Synergistic Φ Earth Mover's Distance Synergistic Information However, computing integrated information for large neurophysiological datasets has been challenging due to both, computational difficulties and limits on do- mains where these measures can be implemented. For instance, many of these measures use the minimum information partition of the network. This involves evaluating a large number of network configurations (more precisely, the Bell number), which makes their computational cost extremely high for large net- works. As for domains of applicability, the measure of [12] has been formu- lated for discrete-state, deterministic, Markovian systems with the maximum entropy distribution. On the other hand, the measure of [14] has been devised to continuous-state, stochastic, non-Markovian systems and in principle, admits dynamics with any empirical distribution (although in practice, it is easier to use assuming Gaussian distributions). The formulation in [14] is based on mutual information, whereas [12] uses a measure based on the Kullback-Leibler diver- gence. Note however, that in some cases the measure of [14] can take negative values and that complicates its interpretation. The Kullback-Leibler based defi- nition computes the information generated during state transitions and remains positive in the regime of stable dynamics. This gives it a natural interpretation as an integrated information measure. Both measures [12], [14] make use of a normalization scheme in their formulations. Normalization inadvertently intro- duces ambiguities in computations. The normalization is actually used for the purpose of determining the partition of the network that minimizes the inte- grated information, but a normalization dependent choice of partition ends up influencing the value and interpretation of Φ. An alternate measure based on the Earth Mover's distance was proposed in [29]. This does away with the normaliza- tion problem (though the current version is not formulated for continuous-state variables). However, the formulation of [29] lies outside the scope of standard information theory and is still difficult for performing computations on large networks. More recently, these issues have been addressed in [9], using a formulation of stochastic integrated information based on the Kullback-Leibler divergence between the conditional multivariate distribution on the set of network states versus the corresponding factorized distribution over its parts, while implement- ing the maximum information partition instead of the minimum information partition. Using this formulation, Φ can be computed for large-scale networks with linear stochastic dynamics, for both, attractor as well as non-stationary states [9] (for network simulations see [2], [3], [10]). This work also demonstrated the first computation of Φ for the resting-state human brain connectome. The connectome network is estimated from cortical white matter tractography data, comprising 998 voxels (nodes) with approximately 28,000 weighted symmetric connections [24]. [9] show that the dynamics and topology of the healthy resting- state brain generates greater information complexity than a (weight-preserving) random rewiring of the same network. Even though this formulation of stochas- tic integrated information was successfully implemented for the human cerebral connectome, a network of 998 nodes and about 28,000 edges, it was limited to linearized dynamics. This is well-defined in the vicinity of attractor states such as the resting-state, however, it would be desirable to extend this formulation to include non-linearities existing in brain dynamics. 3 Empirical Measures Ideally, integrated information was intended as a measure of awareness, one that could account for informational differences between states and also disorders of consciousness. However, as described above, for realistic brain dynamics and physiological data that task has in fact proven difficult. On the other hand, the basic conceptualization of consciousness in terms of integration and differentia- tion of causal information has motivated several empirical measures that seek to classify consciousness-related disorders from patient data. For example, [13] in- vestigated changes in conscious levels using Granger Causality (GC) as a causal connectivity measure. Given two stationary time-series signals, Granger Causal- ity measures the extent to which the past of one assists in predicting the future of the other, over and above the extent to which the past of the latter already predicts its own future [21], thus quantifying causal relations between two sig- naling sources. This was tested using electroencephalographic (EEG) data from subjects undergoing propofol-induced anesthesia, with signals source-localized to the anterior and posterior cingulate cortices. [13] found a significant increases in bidirectional GC in most subjects during loss of consciousness, especially in the beta and gamma frequency ranges. Another useful measure of causal connec- tivity is transfer entropy, which extends Granger causality to the non-Gaussian case. However, so far this has only been implemented on neuronal cultures by [40] and holds future potential as a clinically relevant measure. Yet another measure that has already proven useful as a clinical classifier of conscious lev- els is the Perturbational Complexity Index (PCI), which was introduced by [18] and tested on TMS-evoked potentials measured with EEG. PCI is calculated by perturbing the cortex with transcranial magnetic stimulation (TMS) in or- der to engage distributed interactions in the brain and then compressing the resulting spatiotemporal EEG responses to measure their algorithmic complex- ity, based on the Lempel-Ziv compression. For a given segment of EEG data, the Lempel-Ziv algorithm quantifies complexity by counting the number of dis- tinct patterns in the data. For example, this can be proportional to the size of a computer file after applying a data compression algorithm. Computing the Lempel-Ziv compressibility requires binarizing the time-series data, based either on event-related potentials or with respect to a given threshold. Using PCI, [18] were able to discriminate levels of consciousness during wakefulness, sleep, and anesthesia, as well as in patients who had emerged from coma and recovered a minimal level of consciousness. Later, the Lempel-Ziv complexity was also used by [31] on spontaneous high-density EEG data recorded from subjects under- going propofol-induced anesthesia. Once again, a robust decline in complexity was observed during anesthesia. These are complexity measures based on data compression algorithms. A qualitative comparison between a data compression measure inspired by PCI and Φ was made in [38]. While compression-based measures do seem to capture certain aspects of Φ, the exact relationship be- tween the two is not completely clear. Nonetheless, these empirical measures have been useful for clinical purposes, in terms of broadly discriminating dis- orders of consciousness. Another relevant complexity measure is the weighted symbolic mutual information (wSMI), introduced by [25]. This is a measure of global information sharing across brain areas. It evaluates the extent to which two EEG channels present nonrandom joint fluctuations, suggesting that they share common sources. This is done by first transforming continuous signals into discrete symbols, and subsequently computing the joint probabilities of symbol pairs between two EEG channels. Before computing the symbolic mutual infor- mation between two time-series signals, a weighting is introduced to disregard conjunctions of identical or opposite-sign symbols from the two signal trains as that could potentially arise from common-source artifacts. In [25] wSMI was estimated for 181 EEG recordings from awake but noncommunicating patients diagnosed in various disorders of consciousness (including 143 from patients in vegetative and minimally conscious states). This measure of information sharing was found to systematically increases with consciousness. In particular, it was able to distinguish patients in the vegetative state, minimally conscious state, and fully conscious state. In Table 2 we summarize the above empirical mea- sures along with their domains of application. 4 Discussion The paradigm-shifting proposal that consciousness might be measurable in terms of the information generated by causal dynamics of the brain as a whole, over the sum of its parts, has led to precise quantitative formulations of information- theoretic complexity measures. These measures seek to operationalize the intu- ition that the complexity associated to consciousness arises from simultaneous integration and differentiation of the brain's structural and dynamical hierar- chies. However, progress in this direction has faced practical challenges such Table 2. Empirical complexity measures alongside their tested domains of application. Empirical Measures Tested Application Domains Granger Causality Wakefulness vs propofol-induced anesthesia using EEG Perturbational Complexity Index Wakefulness, sleep, anesthesia, coma & minimal Lempel-Ziv Complexity Weighted Symbolic Mutual Information consciousness using TMS-evoked EEG Wakefulness vs propofol-induced anesthesia using EEG Vegetative, minimally conscious & fully conscious states using EEG as high computational cost upon scaling with network size. This is especially true with regard to realistic neuroimaging or physiological datasets. Even in the approach of [9], where both, the scaling and normalization problem have been solved, the formulation is still applicable only to linear dynamical systems. A possible way to extend this formulation to non-linear systems such as the brain might be to first solve the Fokker-Planck equations for these systems (as prob- ability distributions will no longer remain Gaussian) and subsequently estimate entropies and conditional entropies numerically to compute Φ. Another solution to the problem might be to construct statistical estimators for the covariance matrices from data and then compute Φ. In the meanwhile, for clinical purposes, it has been useful to consider empiri- cal complexity measures, which serve as classifiers that very broadly discriminate states of consciousness, such as between wakefulness and anesthesia or broadly between disorders of consciousness. However, these measures do not strictly cor- respond to integrated information. Some of them are based on signal compres- sion, which does capture differentiation, though not directly integration. So far these methods have been applied on the scale of EEG datasets. One has yet to demonstrate their computational feasibility for larger datasets (which might only be a matter of time though). All in all, bottom-up approaches suggest important features that might help inform or constrain implementations of principle-based approaches. However, the latter are indispensable for ultimately understanding causal aspects of information generation and flow in the brain. This article is intended as a lookup table spanning the landscape of both, theoretically-motivated as well as empirically-based complexity measures used in current consciousness research. Even though, for the purpose of this article, we have treated complexity as a global correlate of consciousness, there are indica- tions that multiple complexity types, based on cognitive and behavioral control, might be important for a more precise classification of various states of con- sciousness [15], [6]. This latter observation alludes to the need for an integrative systems approach to consciousness research, one that is grounded in cognitive architectures and helps understand control mechanisms underlying systems level neural information processing [4]. Acknowledgments. This work has been supported by the European Research Council's CDAC project: "The Role of Consciousness in Adaptive Behavior: A Combined Empirical, Computational and Robot based Approach" (ERC-2013- ADG 341196). References 1. Arsiwalla, X.D., Verschure, P.F.M.J.: Integrated information for large complex net- works. In: The 2013 International Joint Conference on Neural Networks (IJCNN). pp. 1–7 (Aug 2013) 2. Arsiwalla, X.D., Betella, A., Bueno, E.M., Omedas, P., Zucca, R., Verschure, P.F.: The dynamic connectome: A tool for large-scale 3d reconstruction of brain activity in real-time. In: ECMS. pp. 865–869 (2013) 3. Arsiwalla, X.D., Dalmazzo, D., Zucca, R., Betella, A., Brandi, S., Martinez, E., Omedas, P., Verschure, P.: Connectomics to semantomics: Addressing the brain's big data challenge. Procedia Computer Science 53, 48–55 (2015) 4. Arsiwalla, X.D., Herreros, I., Moulin-Frier, C., Sanchez, M., Verschure, P.F.: Is Consciousness a Control Process?, pp. 233–238. IOS Press, Amsterdam (2016) 5. Arsiwalla, X.D., Mediano, P.A., Verschure, P.F.: Spectral modes of network dynam- ics reveal increased informational complexity near criticality. Procedia Computer Science 108, 119–128 (2017) 6. Arsiwalla, X.D., Moulin-Frier, C., Herreros, I., Sanchez-Fibla, M., Verschure, P.F.: The morphospace of consciousness. arXiv preprint arXiv:1705.11190 (2017) 7. Arsiwalla, X.D., Verschure, P.: Computing Information Integration in Brain Net- works, pp. 136–146. Springer International Publishing, Cham, Switzerland (2016) 8. Arsiwalla, X.D., Verschure, P.F.M.J.: High Integrated Information in Complex Networks Near Criticality, pp. 184–191. Springer International Publishing, Cham, Switzerland (2016) 9. Arsiwalla, X.D., Verschure, P.F.: The global dynamical complexity of the human brain network. Applied Network Science 1(1), 16 (2016) 10. Arsiwalla, X.D., Zucca, R., Betella, A., Martinez, E., Dalmazzo, D., Omedas, P., Deco, G., Verschure, P.: Network dynamics with brainx3: A large-scale simulation of the human brain network with real-time interaction. Frontiers in Neuroinfor- matics 9(2) (2015) 11. Ay, N.: Information geometry on complexity and stochastic interaction. Entropy 17(4), 2432–2458 (2015) 12. Balduzzi, D., Tononi, G.: Integrated information in discrete dynamical systems: motivation and theoretical framework. PLoS Comput Biol 4(6), e1000091 (2008) 13. Barrett, A.B., Murphy, M., Bruno, M.A., Noirhomme, Q., Boly, M., Laureys, S., Seth, A.K.: Granger causality analysis of steady-state electroencephalographic sig- nals during propofol-induced anaesthesia. PloS one 7(1), e29072 (2012) 14. Barrett, A.B., Seth, A.K.: Practical measures of integrated information for time- series data. PLoS Comput Biol 7(1), e1001052 (2011) 15. Bayne, T., Hohwy, J., Owen, A.M.: Are there levels of consciousness? Trends in cognitive sciences 20(6), 405–413 (2016) 16. Blumenfeld, H.: Impaired consciousness in epilepsy. The Lancet Neurology 11(9), 814–826 (2012) 17. Bodart, O., Gosseries, O., Wannez, S., Thibaut, A., Annen, J., Boly, M., Rosanova, M., Casali, A.G., Casarotto, S., Tononi, G., et al.: Measures of metabolism and complexity in the brain of patients with disorders of consciousness. NeuroImage: Clinical 14, 354–362 (2017) 18. Casali, A.G., Gosseries, O., Rosanova, M., Boly, M., Sarasso, S., Casali, K.R., Casarotto, S., Bruno, M.A., Laureys, S., Tononi, G., et al.: A theoretically based index of consciousness independent of sensory processing and behavior. Science translational medicine 5(198), 198ra105–198ra105 (2013) 19. Giacino, J.T.: The vegetative and minimally conscious states: consensus-based cri- teria for establishing diagnosis and prognosis. NeuroRehabilitation 19(4), 293–298 (2004) 20. Giacino, J.T., Ashwal, S., Childs, N., Cranford, R., Jennett, B., Katz, D.I., Kelly, J.P., Rosenberg, J.H., Whyte, J., Zafonte, R., et al.: The minimally conscious state definition and diagnostic criteria. Neurology 58(3), 349–353 (2002) 21. Granger, C.W.: Investigating causal relations by econometric models and cross- spectral methods. Econometrica: Journal of the Econometric Society pp. 424–438 (1969) 22. Griffith, V.: A principled infotheoretic\ phi-like measure. arXiv preprint arXiv:1401.0978 (2014) 23. Griffith, V., Koch, C.: Quantifying Synergistic Mutual Information, (2014), 159–190. pp. http://dx.doi.org/10.1007/978-3-642-53734-9_6 Springer Berlin Heidelberg, Berlin, Heidelberg 24. Hagmann, P., Cammoun, L., Gigandet, X., Meuli, R., Honey, C.J., Wedeen, V.J., Sporns, O.: Mapping the Structural Core of Human Cerebral Cortex. PLoS Biology 6(7), 15 (2008) 25. King, J.R., Sitt, J.D., Faugeras, F., Rohaut, B., El Karoui, I., Cohen, L., Nac- cache, L., Dehaene, S.: Information sharing in the brain indexes consciousness in noncommunicative patients. Current Biology 23(19), 1914–1919 (2013) 26. Krohn, S., Ostwald, D.: Computing integrated information. arXiv preprint arXiv:1610.03627 (2016) 27. Laureys, S.: The neural correlate of (un) awareness: lessons from the vegetative state. Trends in cognitive sciences 9(12), 556–559 (2005) 28. Laureys, S., Owen, A.M., Schiff, N.D.: Brain function in coma, vegetative state, and related disorders. The Lancet Neurology 3(9), 537–546 (2004) 29. Oizumi, M., Albantakis, L., Tononi, G.: From the phenomenology to the mech- anisms of consciousness: integrated information theory 3.0. PLoS Comput Biol 10(5), e1003588 (2014) 30. Parton, A., Malhotra, P., Husain, M.: Hemispatial neglect. Journal of Neurology, Neurosurgery & Psychiatry 75(1), 13–21 (2004) 31. Schartner, M., Seth, A., Noirhomme, Q., Boly, M., Bruno, M.A., Laureys, S., Bar- rett, A.: Complexity of multi-dimensional spontaneous eeg decreases during propo- fol induced general anaesthesia. PloS one 10(8), e0133532 (2015) 32. Seth, A.K.: Causal connectivity of evolved neural networks during behavior. Net- work: Computation in Neural Systems 16(1), 35–54 (2005) 33. Tegmark, M.: Improved measures of integrated information. arXiv preprint arXiv:1601.02626 (2016) 34. Tononi, G.: An information integration theory of consciousness. BMC neuroscience 5(1), 42 (2004) 35. Tononi, G.: Integrated information theory of consciousness: an updated account. Arch Ital Biol 150(2-3), 56–90 (2012) 36. Tononi, G., Sporns, O.: Measuring information integration. BMC neuroscience 4(1), 31 (2003) 37. Tononi, G., Sporns, O., Edelman, G.M.: A measure for brain complexity: relating functional segregation and integration in the nervous system. Proceedings of the National Academy of Sciences 91(11), 5033–5037 (1994) 38. Virmani, M., Nagaraj, N.: A compression-complexity measure of integrated infor- mation. arXiv preprint arXiv:1608.08450 (2016) 39. Wennekers, T., Ay, N.: Stochastic interaction in associative nets. Neurocomputing 65, 387–392 (2005) 40. Wibral, M., Vicente, R., Lindner, M.: Transfer Entropy in Neuroscience, pp. 3–36. Springer Berlin Heidelberg, Berlin, Heidelberg (2014)
1608.06548
3
1608
2016-10-27T15:31:08
Grand Challenges for Global Brain Sciences
[ "q-bio.NC" ]
The next grand challenges for society and science are in the brain sciences. A collection of 60+ scientists from around the world, together with 10+ observers from national, private, and foundations, spent two days together discussing the top challenges that we could solve as a global community in the next decade. We eventually settled on three challenges, spanning anatomy, physiology, and medicine. Addressing all three challenges requires novel computational infrastructure. The group proposed the advent of The International Brain Station (TIBS), to address these challenges, and launch brain sciences to the next level of understanding.
q-bio.NC
q-bio
Grand Challenges for Global Brain Sciences Global Brain Workshop 2016 Attendees* The next grand challenges for science and society are in the brain sciences. A collection of 60+ scientists from around the world, together with 15+ observers from national, private, and foundations, spent two days together discussing the top challenges that we could solve as a global community in the next decade. We settled on three challenges, spanning anatomy, physiology, and medicine. Addressing all three challenges requires novel computational infrastructure. The group proposed the creation of The International Brain Station (TIBS), to address these challenges, and launch brain sciences to the next level of understanding. Understanding the brain and curing its diseases are among the most exciting challenges of our time. Consequently, national, transnational, and private parties are investing billions of dollars (USD). ​To efficiently join forces, ​Global Brain Workshop 2016 was hosted at Johns Hopkins University's Kavli Neuroscience Discovery Institute on April 7-8. A second workshop, ​Open Data Ecosystem in Neuroscience took place July 25-26 in Washington, DC to continue the discussion specifically about computational challenges and opportunities. A third conference, ​Coordinating took place in New York City on September 19th in association with the Global Brain Projects, ​ United Nations General Assembly. So vast are both the challenges and the opportunities that global coordination is crucial. To find ways of synergistically studying the brain, the kick-off workshop welcomed over 60 scientists, representing 12 different countries and a wide range of subdisciplines. They were joined by 15 observers from various national and international funding organizations. Participants were engaged weeks before the conference and charged with coming up with ambitious projects that are both feasible and internationally inclusive, on par with the International Space Station (i.e., worthy of a global, decade-long effort). Over the course of 36 hours, scientists discussed, debated, and gathered feedback, ultimately proposing several "grand challenges for global brain sciences" that were refined by working groups. The workshop was covered in a ​media piece​ in ​Science The group began with 60+ ideas, each forged independently by one of the scientific participants. Each participant proposed a unique challenge that was designed to meet the following desiderata: April 15, 2016. ​ 1. Significant : it will yield tangible societal, economic, and medical benefits to the world. 2. Feasible : it can achieve major milestones within 10 years given existing funding 3. opportunities. : nations throughout the world can meaningfully contribute to and benefit from Inclusive each challenge, and the collection of challenges are collectively scientifically diverse. the proposed ideas were similar to one another and others were Interestingly, a lot of complementary. This allowed the group to converge on three grand challenges for global brain sciences, each depending on a common universal resource. Challenge 1: What makes our brains unique? Both within and across species, brain structure is known to exhibit significant variability across many orders of magnitude in scale-​including anatomy, biochemistry, connectivity, development, and gene expression (ABCDE). It remains mysterious how and why the nervous system tightly regulates certain properties, while allowing others to vary. Understanding the design principles governing variability may hold the key to understanding intelligence and subjective experience, as well as the influence of variability on health and function. This grand challenge is a global project to coordinate the construction of comprehensive multiscale maps of the ABCDE's of multiple brains from multiple species using multiple cognitive and mental health disease models​. Within a decade, we expect to have addressed this challenge in brains including but not limited to Drosophila, Zebrafish, Mouse, and Marmoset, and to have developed tools to conduct massive neurocartographic analyses. The result will be a state-of-the-art "Virtual NeuroZoo" with fully annotated data and analytic tools for analysis and discovery. This virtual NeuroZoo can be utilized by neuroscientists and citizens alike, both as a reference and for educational materials. By incorporating disease models, we explicitly link this challenge with the third challenge. Challenge 2: How does the brain solve complex computational problems? Brains remain the most computationally advanced machines for a large array of cognitive tasks-whether navigating hazardous terrain, translating languages, conducting surgery, or recognizing emotional states-despite the fact that modern computers can utilize millions of training samples, megawatts of power, and tons of hardware. While the ABCDEs establish the "wetware" upon which our brains can solve such computations, to understand the mechanisms we need to measure, manipulate, and model neural activity simultaneously across many resolutions and scales-including wearables, embedded sensors, and spatiotemporal actuators-while animals are exhibiting complex ecological behaviors in naturalistic environments. This grand challenge is a global project to investigate a single naturalistic behavior that is ecologically relevant across phylogenies, such as foraging, and measure brain and body properties across spatial, temporal, and genetic scales​. The challenge differs from previous efforts in three key ways. First, it requires studying animals in ​complex ​and naturalistic it requires ​coordinated attacks at many different scales by many environments. Second, investigators while the animals are performing the same complex behaviors. We different envision groups of 20-30 investigators all operating together to share data and experimental design. Third, the richness of the mental repertoire of cognition suggests that deciphering its codes will require ​many parallel investigations to uncover different facets of brain function. These experiments in turn will produce multiscale models of neural systems with the potential to accomplish computational tasks that no current computer system can perform. ​Mechanistic studies, guided by theoretical models, will help to ask how perturbations of those systems lead to aberrant function, linking this challenge with the next one. financial costs, and loss of productivity. Despite a growing awareness of Challenge 3: How can we augment clinical decision-making to prevent disease and restore brain function? illnesses levy enormous burdens upon humanity: impairment, Psychiatric and neurological suffering, the challenges, clinicians consistently battle the lack of objective tests to guide clinical treatments, prognosis). Compounding these decision-making (e.g., diagnosis, selection of limitations are societal stigmas regarding mental illness that increase the suffering of patients and their families. The ABCDEs of neurobiological variability, when coupled with multiscale mechanistic models of cognition, will provide new approaches to neurobiologically-informed clinical decision making. This grand challenge is a global project to transform clinical decision-making via incorporating neural mechanisms of dysfunction. This will require collecting, organizing and analyzing human and non-human anatomical and functional data. These data, and the tools developed to explore and discover novel treatment therapies, will be the foundation upon which the next decades of experiments and clinical decisions will be based. The distributed and multimodal nature of these datasets further motivate the need for an all-purpose computational platform, upon which models of disease can be developed, deployed, tested, and refined. A Universal Resource All three of the grand challenges for global brain sciences represent severe methodological challenges, both technological and computational. The technological developments required for each of the challenges are non-overlapping. In contrast, regardless of the nature of the scientific questions or data modalities involved, each project will require computational capabilities including collecting, storing, exploring, analyzing, modeling, and discovering data. Although neuroscience has developed a large number of computational tools to deal with existing datasets, the datasets proposed here bring with them a whole suite of new challenges. This resource would be a comprehensive computational platform, deployed in the cloud, that will provide web services for all the current "pain points" in daily neuroscience practice associated with big data. This resource will realize a new era of brain sciences, one in which the bottlenecks to discovery transition away from data collection and processing to data enriching exploring, and modeling. While science has always benefitted from standing on the shoulders of . Today, essentially every giants, this will enable science to stand on the shoulders of everyone practicing neuroscientist's productivity is limited due to computational resources, access to data or algorithms, or struggling with determining which data and algorithms are best suited to answer the most pressing questions of our generation. This resource will create a future where those limitations will feel as archaic as fitting the data with paper and pencil feels today. For further details, see an upcoming NeuroView called "To the Cloud! A Grassroots Proposal to Accelerate Brain Science Discovery". Societal Considerations Each nation affords different opportunities and restrictions, owing to ethical, policy, and cultural considerations. Because these grand challenges are inherently inclusive, manifesting them will require understanding and mitigating issues that arise in cross-cultural endeavors. Indeed, addressing the vast diversity of partnerships in such an endeavor is a challenge in itself. We therefore recommend the following. First, form a ​cultural sensitivity ​committee to consider and investigate potentially sensitive issues. Second, bolstered by their research, establish cross-cultural collaboration education materials​, including written guidelines and videos, which will be recommended to all participating scientists. Third, to deepen the understanding of transnational collaborations, develop ​trainee exchange programs in which participating trainees will spend six months to a year working and training in a foreign country. This will also facilitate cross-cultural knowledge dissemination and fertilization. Fourth, require ​frequent assessments to ensure maintenance of cultural sensitivities. These assessments will feedback into the educational material and be used to modify the exchange programs. Next Steps Crucial to the success of this endeavor is a sequence of actionable steps that the community can follow. Because we are not proposing any additional funding, realizing the eventual goals of these grand challenges will rely on marshalling existing funds. Due to the incoming leadership changes, both on national and transnational levels, quick action is of the essence. Therefore, we have taken the following steps: We have created a webpage, ​http://brainx.io​, containing a bibliography of reports that resulted from this conference, as well as a list of all scientific participants and observers who attended the original brainstorming meeting leading to this document. We will also be monitoring comments on ​https://neurostars.org/ ​w​i​t​h ​t​h​e ​t​a​g "neurostorm​" for further discussion. Finally, we will have an outpost at the NeuroData booth (#4126) at the SfN meeting in San Diego to discuss these issues further. We encourage anybody who feels inspired by this document to join the discussion, engage, and get in touch with funders and other scientists with your ideas. Acknowledgements National Science Foundation (​1637376) and the Kavli Foundation. *Global Brain Workshop 2016 Attendees Joshua T. Vogelstein​1,27,28,29,30,31 ● Katrin Amunts​7,8 ● Andreas Andreou​30 ● Dora Angelaki​32 ● Giorgio A. Ascoli​33 ● Cori Bargmann​34 ● Randal Burns​28, 29 ● Corrado Cali​11 ● Frances Chance​35 ● George Church​36 ● Hollis Cline​37 ● Todd Coleman​38 ● Stephanie de La Rochefoucauld​39 ● ● Winfried Denk​40 Ana Belén Elgoyhen​41 ● ● Ralph Etienne Cummings​42 Alan Evans​5 ● Kenneth Harris​43 ● ● Michael Hausser​3 ● Sean Hill​9 ● Samuel Inverso​44 Chad Jackson​45 ● Viren Jain​46 ● ● Rob Kass​47 ● Bobby Kasthuri​13 Adam Kepecs​15 ● Gregory Kiar​1, 27 ● ● Dean M. Kleissas​25 ● Konrad Kording​6 ● Sandhya P. Koushika​10 John Krakauer​48 ● Story Landis​49 ● ● Jeff Layton​50 ● Qingming Luo​51 Adam Marblestone​52 ● David Markowitz​26 ● ● Justin McArthur​53 ● Brett Mensh​2,4 ● Michael P. Milham​19 Partha Mitra​15 ● Pedja Neskovic​54 ● ● Miguel Nicolelis​55 ● Richard O'Brien​56 Aude Oliva​57 ● Gergo Orban​58 ● ● Hanchuan Peng​14 Eric Perlman​27 ● ● Marina Picciotto​59 ● Mu-Ming Poo​17 ● ● ● ● ● ● Jean-Baptiste Poline​18 Alexandre Pouget​60 Sridhar Raghavachari​61 Jane Roskams​14 Alyssa Picchini Schaffer​20 Terry Sejnowski​62 Friedrich T. Sommer​63 ● Nelson Spruston​4 ● Larry Swanson​64 ● Arthur Toga​65 ● R. Jacob Vogelstein​26 ● Anthony Zador​15 ● ● Richard Huganir​30,31 ● Michael I. Miller​1,27,31 1. Department of Biomedical Engineering, Institute for Computational Medicine, Johns Hopkins University, Baltimore, MD, USA Janelia Research Campus, Howard Hughes Medical Institute, Ashburn, VA, USA 2. Optimize Science, Mill Valley, CA USA; UCSF Kavli Institute for Fundamental Neuroscience, San Francisco, CA, USA 3. Department of Physiology, University College London, London, UK 4. 5. Montreal Neurological Institute, McGill University,,Montreal, Quebec, Canada 6. Physical Medicine and Rehabilitation, Physiology, and Applied Mathematics, and Biomedical Engineering, Northwestern University, Chicago, IL, USA Institute for Neuroscience and Medicine, INM-1, Research Centre Juelich, Germany, C. and O. Vogt Institute for Brain Research, Forschungszentrum Jülich; University Hospital Duesseldorf, University Duesseldorf, Germany 7. Blue Brain Project, EPFL, Campus Biotech, Geneva, Switzerland 8. Human Brain Project, EPFL, Geneva, Switzerland 9. 10. Department of Biological Sciences, Tata Institute of Fundamental Research, Navy Nagar, Colaba, Mumbai, India 11. Biological and Environmental Science and Engineering, KAUST,Thuwal, 23955-6900 Saudi Arabia 12. Cuban Neuroscience Center, 190 e / 25 and 27, Cubanacan, Playa. Havana. CP 11600; University of Electronic Science and Technology of China, Shahe Campus:No.4, Section 2, North Jianshe Road, 610054, Chengdu, Sichuan, P.R.China 13. Argonne National Laboratory, Argonne, IL, USA 14. Allen Institute for Brain Science, Seattle, WA, USA 15. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, USA 16. Department of Psychological and Brain Sciences, Dartmouth College, Hanover, NH, USA 17. Institute of Neuroscience, CAS Center for Brain Science, 320 Yue Yang Road Shanghai, 200031 P.R.China; Intelligence Technology, Chinese Academy of Sciences, 319 Yueyang Road, Shanghai 200031, P.R.China 18. Henry H. Wheeler Jr. Brain Imaging Center, Helen Wills Neuroscience Institute, 188 Li Ka Shing Center for Biomedical and Health Sciences, Henry H. Wheeler, Jr. Brain Imaging Center, Suite B107, University of California, Berkeley, CA 94720, USA 19. Center for the Developing Brain, Child Mind Institute, New York, NY; Nathan S. Kline Institute for Psychiatric Research, Orangeburg, NY, USA 20. Simons Collaboration on the Global Brain, Simons Foundation, New York, NY, USA 21. 22. Department of Physiology, Keio University School of Medicine, 35 Shinanomachi, Shinjuku-ku, Tokyo,, Japan; RIKEN Israel Brain Technologies, Hakfar Hayarok, Ramat Hasharon, Israel Brain Science Institute, Laboratory for Marmoset Neural Architecture, 2-1 Hirosawa, Wako, Saitama, Japan 23. Mind Research Network, Department of Electrical and Computer Engineering, University of New Mexico, Albuquerque, NM, USA Intelligence Advanced Research Projects Activity (IARPA), Maryland Square Research Park, Riverdale Park, MD, USA Institute for Data Intensive Engineering and Sciences, Johns Hopkins University, Baltimore,MD, USA 24. The Kavli Foundation, Oxnard, CA, USA 25. Johns Hopkins University Applied Physics Laboratory, Laurel, MD, USA 26. 27. Center for Imaging Science, Johns Hopkins University, Baltimore, MD, USA 28. Department of Computer Science, Johns Hopkins University, Baltimore, MD, USA 29. 30. Department of Neuroscience, Johns Hopkins University, Baltimore, MD, USA 31. Kavli Neuroscience Discovery Institute, Johns Hopkins University, Baltimore, MD, USA 32. Dept of Neuroscience, Baylor College of Medicine, Houston, TX, USA 33. Dept of Molecular Neuroscience, George Mason University, Fairfax, VA, USA 34. Howard Hughes Medical Institute, Rockefeller University, New York, NY, USA 35. Sandia National Laboratories, Albuquerque, NM, USA 36. Harvard Medical School, Harvard University, Boston, MA, USA 37. Department of Molecular and Cellular Neuroscience, The Scripps Research Institute, La Jolla, CA, USA 38. Department of Bioengineering, University of California, San Diego, CA, USA International Brain Research Organization (IBRO), Paris, France 39. 40. Max Planck Institute of Neurobiology, Martinsried, Germany 41. Molecular Biology and Genetic Engineer Institute, CONICET, Argentina 42. Department on Electrical & Computer Engineering, Johns Hopkins University, Baltimore, MD, USA 43. Department of Quantative Neuroscience, University College London, London, England 44. Department of Genetics, Harvard University, Boston, MA, USA 45. U.S. Department of State, Washington D.C., USA 46. Google, Mountain View, CA, USA 47. Department of Statistics, Carnegie Mellon University, Pittsburgh, PA, USA 48. Department of Neurology, Johns Hopkins Hospital, Baltimore, MD, USA 49. National Institute of Neurological Disorders and Stroke (NINDS), National Institute of Health, Bethesda, MD, USA 50. Amazon Web Services, Atlanta, GA, USA 51. Britton Chance Center for Biomedical Photonics, Wuhan National Laboratory for Optoelectronics, Huazhong University of Science & Technology, Wuhan, 430074, China 52. MIT Media Lab, Cambridge, MA, USA 53. Department of Neurology, Johns Hopkins Hospital, Baltimore, MD, USA 54. Office of Naval Research, Arlington, VA, USA 55. Duke Institute for Brain Sciences, Duke University, Durham, NC, USA 56. Department of Neurology, Duke University School of Medicine, Durham, NC, USA 57. National Science Foundation, Arlington, VA, USA 58. Department of Theoretical Physics, MTA Wigner Research Centre for Physics, Budapest, Hungary 59. Department of Psychiatry, Yale University, New Haven, CT, USA 60. Neuroscience Center, University of Geneva, Geneva, Swizterland 61. National Science Foundation, Arlington, VA, USA 62. Salk Institute for Biological Studies, La Jolla, CA, USA 63. Redwood Center for Theoretical Neuroscience, University of California, Berkley, CA, USA 64. Department of Biological Sciences, University of Southern California, Los Angeles, CA, USA 65. Institute for Neuroimaging and Informatics, University of Southern California, Los Angeles, CA, USA
1604.07457
1
1604
2016-04-25T22:01:50
Modeling the Contribution of Central Versus Peripheral Vision in Scene, Object, and Face Recognition
[ "q-bio.NC", "cs.CV" ]
It is commonly believed that the central visual field is important for recognizing objects and faces, and the peripheral region is useful for scene recognition. However, the relative importance of central versus peripheral information for object, scene, and face recognition is unclear. In a behavioral study, Larson and Loschky (2009) investigated this question by measuring the scene recognition accuracy as a function of visual angle, and demonstrated that peripheral vision was indeed more useful in recognizing scenes than central vision. In this work, we modeled and replicated the result of Larson and Loschky (2009), using deep convolutional neural networks. Having fit the data for scenes, we used the model to predict future data for large-scale scene recognition as well as for objects and faces. Our results suggest that the relative order of importance of using central visual field information is face recognition>object recognition>scene recognition, and vice-versa for peripheral information.
q-bio.NC
q-bio
Modeling the Contribution of Central Versus Peripheral Vision in Scene, Object, and Face Recognition Panqu Wang ([email protected]) Department of Electrical and Engineering, University of California San Diego 9500 Gilman Dr 0407, La Jolla, CA 92093 USA Garrison W. Cottrell ([email protected]) Department of Computer Science and Engineering, University of California San Diego 9500 Gilman Dr 0404, La Jolla, CA 92093 USA 6 1 0 2 r p A 5 2 ] . C N o i b - q [ 1 v 7 5 4 7 0 . 4 0 6 1 : v i X r a Abstract It is commonly believed that the central visual field (fovea and parafovea) is important for recognizing objects and faces, and the peripheral region is useful for scene recognition. However, the relative importance of central versus peripheral informa- tion for object, scene, and face recognition is unclear. Larson and Loschky (2009) investigated this question in the context of scene processing using experimental conditions where a cir- cular region only reveals the central visual field and blocks peripheral information ("Window"), and in a "Scotoma" con- dition, where only the peripheral region is available. They measured the scene recognition accuracy as a function of vi- sual angle, and demonstrated that peripheral vision was indeed more useful in recognizing scenes than central vision in terms of achieving maximum recognition accuracy. In this work, we modeled and replicated the result of Larson and Loschky (2009), using deep convolutional neural networks (CNNs). Having fit the data for scenes, we used the model to predict future data for large-scale scene recognition as well as for ob- jects and faces. Our results suggest that the relative order of importance of using central visual field information is face recognition>object recognition>scene recognition, and vice- versa for peripheral information. Furthermore, our results pre- dict that central information is more efficient than peripheral information on a per-pixel basis across all categories, which is consistent with Larson and Loschky's data. Keywords: face recognition; object recognition; scene recog- nition; central and peripheral vision; deep neural networks Introduction Viewing a real-world scene occupies the entire visual field, but the visual resolution across the visual field varies. The fovea, a small region in the center of the visual field that sub- tends approximately 1◦ of visual angle (Polyak, 1941), per- ceives the highest visual resolution of 20 to 45 cycles/degree (cpd) (Loschky, McConkie, Yang, & Miller, 2005). The parafovea has a slightly lower visual resolution and extends to about 4-5◦ eccentricity, where the highest density of rods is found (Wandell, 1995). Beyond the parafovea is generally considered to be peripheral vision (Holmes, Cohen, Haith, & Morrison, 1977), which receives the lowest visual resolution. Due to the high density and small receptive field of retinal receptors,the central (foveal and parafoveal) vision encodes information of higher spatial frequency and more detail; pe- ripheral vision, on the contrary, encodes coarser and lower spatial frequency information. This retinotopic representation of the visual field is mapped to visual cortical areas through a log-polar representation. Recent studies have shown that orderly central and periph- eral representations can be found not only in low-level to mid-level visual areas (V1-V4), but also in higher-level re- gions, where perception and recognition for faces or scenes is engaged (Malach, Levy, & Hasson, 2002; Grill-Spector & Malach, 2004). More specifically, Malach et al. (2002) pro- posed that the need for visual resolution is a crucial factor in organizing object areas in higher-level visual cortex: ob- ject recognition that depends more on fine detail is associated with central-biased representations, such as faces and words; object recognition that depends more on large-scale integra- tion is associated with peripheral-biased representations, such as buildings and scenes. This hypothesis is supported by fMRI evidence, which shows that the brain areas that are more activated for faces (FFA; Kanwisher, McDermott, and Chun (1997)) and words (VWFA; McCandliss, Cohen, and Dehaene (2003)) sit in the eccentricity band expanded by central visual-field bias, whereas buildings and scenes (PPA; Epstein, Harris, Stanley, and Kanwisher (1999)) are asso- ciated with peripheral bias. More recent studies even sug- gest that the central-biased pathway for recognizing faces and peripheral-biased pathway for recognizing scenes are segre- gated by mid-fusiform sulcus (MFS) to enable fast parallel processing (Gomez et al., 2015). In the domain of behavioral research, studies have shown that object perception performance is the best around 1◦-2◦ of fixation point and drops rapidly as eccentricity increases (Henderson & Hollingworth, 1999; Nelson & Loftus, 1980). For scene recognition, Larson and Loschky (2009) used a "Window" and "Scotoma" design (see Figure 1), to test the contributions of central versus peripheral vision to scene recognition. The Window condition (top rows of the right- hand columns of Figure 1) presents central information at various visual angles to the subjects, while the Scotoma con- dition (second row on the right) blocks it. Using images from 10 categories, subjects were required to verify the category in each condition. The recognition accuracy as a function of visual angle is shown in Figure 2. They found that foveal vision is not accurate for scene perception, while peripheral vision is, despite its much lower resolution. However, they also found that central vision is more efficient, in the sense that less area is needed to achieve equal accuracy. The visual area is equal at 10.8◦, and the crossover point, where central vision starts to perform better than peripheral, is to the left of that point. Despite the common belief that central vision is important for face and object recognition, and peripheral vision is im- portant for scene perception shown in studies above, a careful examination of the contribution of central versus peripheral vision in object, scene, and face recognition is needed. In this work, we modeled the experiment of Larson and Loschky (2009) using deep convolutional neural networks. Further- more, we extended the modeling work to a greater range of stimuli, and answer the following questions: How does the model perform as the number of scene categories is scaled up? Besides scenes, can the model predict the importance of central vision versus peripheral information in object and face recognition? What is the result compared to scenes? In the following, we show that our modeling results match the observations of Larson and Loschky (2009), and that it scales up to over 200 scene categories. By running a sim- ilar analysis for large-scale object and face recognition, our model predicts that central vision is very important for face recognition, important for object recognition, and less impor- tant for scene recognition. Peripheral vision, however, serves an important role for scene recognition, but is less impor- tant for recognizing objects and faces. Furthermore, across all conditions we tried, central vision is more efficient than peripheral vision on a per-pixel basis (when equal areas are presented), which is consistent with the result of Larson and Loschky (2009). Method Image Preprocessing To create foveated images, we preprocessed the images using the Space Variant Imaging System1. To mimic human vision, we set the parameter that specifies the eccentricity at which resolution drops to half of the fovea to 2.3◦. Example images and their preprocessed retinal versions are shown in the first and second columns of Figure 1. As in the experiments of Larson and Loschky (2009), we used the Window and Scotoma paradigms as specified by van Diepen, Wampers, and dYdewalle (1998) to process the input stimulus. The idea of both paradigms is to evaluate the value of missing information - if the missing information is needed, then the perception process may be disrupted and recognition performance may drop; if the missing information is not nec- essary, then the processing remains normal. Input images in our experiments are 256× 256 pixels, and we assume that corresponds to 27◦ × 27◦ of visual angle, the number in (Larson & Loschky, 2009). In (Larson & Loschky, 2009), they used four sets of radius conditions for Windows and Scotomas: 1◦ represents the presence or ab- sence of foveal vision; 5◦ represents the presence or absence of central vision; 10.8◦ presents equal viewable area inside the Windows or outside the Scotomas; 13.6◦ presents more viewable area in the Windows than the Scotomas. In order make the prediction of the model more accurate, we added five additional radius conditions in all of our experiments: 1http://svi.cps.utexas.edu/software.shtml Figure 1: Examples of images used in our experiment. First column: original images. Second column: foveated images. Third to last column: images processed through "Window" and "Scotoma" conditions with different radii in degrees of visual angle. 3◦,7◦,9◦,12◦, and 16◦. The example Window and Scotoma images are shown in Figure 1. Deep Convolutional Neural Networks (CNNs) Deep CNNs are neural networks with many layers that stack computations in a hierarchical way, repeatedly performing: 1) 2-dimensional convolutions over the stimulus generated from previous layers using learned filters, which are con- nected locally to a small subregion of the visual field; 2) a pooling operation on local regions of the feature maps ob- tained from convolution operation, which is used to reduce the dimensionality and gain translational invariance; 3) non- linearities to the upstream response, which is used to gener- ate more discriminative features useful for the task. As layers go higher, the receptive fields of filters are generally larger, and the learned features go from low-level (edges, contours) to high-level object-related representations (object parts and shapes) (Zeiler & Fergus, 2014). Several fully-connected lay- ers are usually added on top of these computations to learn more abstract and task-related features. We used deep CNNs in our experiments for two reasons. First, deep CNNs are the best models in computer vision: they achieve the state-of-the-art performance on many large- scale computer vision tasks, such as image classification (Krizhevsky, Sutskever, & Hinton, 2012; He, Zhang, Ren, & Sun, 2015), object detection (Ren, He, Girshick, & Sun, 2015), and scene recognition (Zhou, Lapedriza, Xiao, Tor- ralba, & Oliva, 2014). Thus, the models should achieve de- cent performance in our experiments. Smaller networks or other algorithms are not competent for our tasks. Second, deep CNNs have been shown to be the best models of the vi- sual cortex: they are able to explain a variety of neural data in human and monkey IT (Yamins et al., 2014; Guc¸lu & van Gerven, 2015; Wang, Malave, & Cipollini, 2015). As a result, it is natural to use them in our work modeling a behavioral study related to human vision. Experiments In this section, we first describe our model of the behavioral study of Larson and Loschky (2009). We then introduce the experiment for measuring the contribution of central versus peripheral vision for large-scale scene, object, and face recog- nition tasks. Modeling Larson and Loschky (2009) In Larson and Loschky (2009), scene recognition accuracy was measured across 100 human subjects on 10 categories: Beach, Desert, Forest, Mountain, River, Farm, Home, Mar- ket, Pool, and Street. For each trial in the Windows and Sco- tomas conditions, subjects were first presented a scene im- age, and then were asked to press "yes" or "no" for the cue (category name) presented on the screen. Their experimental result is summarized in Figure 2. They showed that central vision (5◦ window condition) performs less well than periph- eral vision in terms of getting maximum recognition perfor- mance. They further demonstrated the peripheral advantage is due to more viewing areas in the Scotomas conditions, and central vision is more privileged when given equal viewable areas (10.8◦). We obtained the stimuli of the above 10 categories from the Places205 database (Zhou et al., 2014), which contains 205 scene categories and 2.5 million images. All input stim- uli were preprocessed using the retina model described in the above section. As 10 categories is small and can easily lead to overfitting problems in training deep CNNs, we trained our recognition model by performing fine-tuning (or transfer learning) based on pretrained models. The model pretrained on the Places205 database can be treated as a mature scene recognition pathway, and fine-tuning can be thought as addi- tional training for the task. To investigate whether different network architectures, especially depth, have different impact on the modeling result, we applied three different pre-trained models, namely: 1. AlexNet (Krizhevsky et al., 2012): A network with 5 con- volutional layers and 3 fully connected layers, about 60 million trainable parameters. Achieved 81.10% top-5 ac- curacy on the Places205 validation set. 2. VGG-16 (Simonyan & Zisserman, 2014): A network with 13 convolutional layers and 3 fully connected layers, about 138 million trainable parameters. Achieved 85.41% top-5 accuracy on the Places205 validation set. 3. GoogLeNet (Szegedy et al., 2015): A network with 21 convolutional layers and 1 fully connected layer, about 6.8 million trainable parameters. Achieved 87.70% top-5 ac- curacy on the Places205 validation set. For all models, the fine-tuning process starts by keeping the weights except for the last fully connected layer intact, and Figure 2: Results for scene recognition accuracy as a func- tion of viewing condition (Windows (w) and Scotomas (s)) and visual angle. Left: result of Larson and Loschky (2009). Right: our modeling result. initializing the weights of the last layer to be random with zero mean and unit variance. To be compatible with the "yes" or "no" condition in the behavioral experiment, we replaced the last layer in the networks with a single logistic unit, and trained the networks for each of the 10 object categories sep- arately, using half of the training images from the target cat- egory and half randomly selected from all other 9 categories. As the last layer needs more learning, we set the learning rate of the last layer to 0.001, and all previous layers to 1e−4. The training set of the 10 scene categories contains a total num- ber of 129,210 full resolution images, and we trained all net- works using minibatch stochastic gradient descent with batch size from 32 to 256, using the Caffe deep learning framework (Jia et al., 2014) on NVIDIA Titan Black 6GB GPUs. All networks were trained for a maximum number of 24,000 it- erations to ensure convergence. Each test set contains 200 images (100 from target category and 100 from all other cat- egories), and the label distribution is the same as the training set.. Test images were preprocessed to meet each of the Win- dows and Scotomas condition. We tested the performance of the fine-tuned models on all conditions by reporting the mean classification accuracy, which is shown in Figure 2. From Figure 2, we can clearly see our result for all three models qualitatively matches the result of Larson and Loschky (2009). First, for Window and Scotoma conditions, an increasing radius of visual angle (x axis) yields a mono- tonic increase or decrease in classification accuracy (y axis). The sharper increase from 1◦ to 5◦ in the behavioral study may be due to the higher efficiency of human central vision. Second, we replicated the fact that central vision (less than 5◦) is less useful than peripheral vision in terms achieving the best scene recognition performance. Third, however, when using equal viewable areas (10.8◦), central vision performs better than peripheral, exhibiting higher efficiency. Fourth, the critical radius (the crossover point where the two con- ditions produce equal performance, see Figure 2b) is 8.26◦ (averaged across all models), which is within the 8.22◦-9.24◦ range reported by Larson and Loschky (2009). This suggests our models are quite plausible. Figure 3: Results for large-scale scene recognition accuracy as a function of viewing condition (Windows (w) and Sco- tomas (s)) and visual angle. Softmax output is used instead of logistic unit, so chance is 0.005. Left: experiment using original images. Right: experiment using foveated images. Figure 4: Results for large-scale object recognition accuracy as a function of viewing condition (Windows (w) and Sco- tomas (s)) and visual angle. Softmax output is used instead of logistic unit, so chance is 0.001. Left: experiment with original images. Right: experiment with foveated images. When comparing the performance across the three models we use, we cannot find a notable difference in terms of perfor- mance, though GoogLeNet usually performs slightly better, indicating that depth of processing might be the key factor in obtaining better performance. Large-Scale Scene, Object, and Face Recognition The above modeling work is based on a scene recognition task using 10 categories. In real life, however, there are a much larger number of scene categories. Beyond scenes, gen- eral object recognition and face recognition are the two most important recognition tasks that are performed regularly. The relative importance of central versus peripheral vision among the three categories needs to be examined carefully. Using a similar modeling approach, we describe our findings in large- scale scene, object, and face recognition in the sections below. Scene Recognition We used all 205 categories in the Places205 dataset. The trained models of AlexNet, VGG-16, and GoogLeNet are deployed to examine the recognition ac- curacy on the Place205 validation set, which contains 20,500 images, in all Windows and Scotoma conditions. In addition, we tested the models using images both processed and un- processed by the retina model to examine the generalization power of the learned features. The result is shown in Figure 3. From Figure 3, we can see the general trend that we ob- served in Figure 2 still holds: peripheral vision is more im- portant than central vision, but central vision is more efficient. All models behave similarly. However, we can see the perfor- mance on images preprocessed through the retina model is inferior. Apparently, since there are many more categories in this experiment, the foveation has more of an effect. Recall that the models are trained using images with full resolution; missing the peripheral information may the cause learned fea- tures to imperfectly generalize. Object Recognition We ran our object recognition experi- ment on the ILSVRC 2012 dataset (Russakovsky et al., 2015), which contains 1000 object categories and over 1.2 million training images. We used the pretrained models of AlexNet, VGG-16, and GoogLeNet, which achieve top-5 accuracy of 80.13%, 88.44%, and 89.00%, respectively, on the ILSVRC 2012 validation set. Similar to scene recognition, we tested all models under all Windows and Scotoma conditions, us- ing original and foveated images. The results are shown in Figure 4. At the first glance of looking at Figure 4, we may draw the conclusion that the result is the same as scene recognition: central vision is still more important than peripheral vision. However, when we compare the scene and object recogni- tion results (shown in Figure 5), we can clearly see that cen- tral information in object recognition is more important than that in scene recognition: the accuracy of the Scotoma con- ditions drops much faster for object recognition than scene recognition as visual angle increases from 1◦ to 7◦, suggest- ing that losing central vision causes a greater impairment for object recognition performance. This is consistent with our knowledge that central vision plays a more important role in object recognition than scenes, as there are more high spa- tial frequency details in objects than scenes. Another find- ing from this experiment is that AlexNet (8 layers) performs much worse than VGG-16 (16 layers) and GoogLeNet (23 layers), suggesting that depth is important to produce good performance. Face Recognition We performed the face recognition ex- periment on the Labeled Faces in the Wild (LFW) dataset (Huang, Ramesh, Berg, & Learned-Miller, 2007), which con- tains 13,233 labeled images from 5,749 individuals. As there is only 1 image for some identities, researchers usually pre- train their network on larger datasets (not publicly available) and test their models on the LFW dataset. In this experiment, we tested three pretrained models, namely Lighten-A (10 lay- ers; (Wu, He, & Sun, 2015)), Lighten-B (16 layers), and VGG-Face (16 layers;(Parkhi, Vedaldi, & Zisserman, 2015)), Figure 5: Comparison results for scene and object recognition using the VGG-16 model. Losing central vision decreases performance for object recognition more quickly than scene recognition. Left: original images. Right: foveated images. on the face verification task for the LFW dataset, where they achieve accuracy of 90.33%, 92.37%, and 96.23%, respec- tively. Face images were preprocessed so that they occupy the entire visual field (Figure 1). Same as the previous ex- periments, we tested all models using Windows and Scotoma conditions, with original and foveated images. Results are shown in Figure 6. We see very different performance in Figure 6 compared to object and scene recognition. First, central information is obviously much more important than peripheral informa- tion for face recognition, given the accuracy at 5◦ is much higher for the Window condition than the Scotoma condition for Lighten models, and very similar with each other for the VGG model. This is consistent with our intuition that face recognition is a fine-grained discrimination process. Second, the Window performance grows much more slowly after 7◦, suggesting the more peripheral region provides little addi- tional information for recognizing faces, unlike objects and scenes, which needs lots of peripheral information to obtain the maximal accuracy. Third, the foveated images produce nearly identical results as the original image, demonstrating that face recognition only involves central vision, and the blurred peripheral vision is not needed. Finally, as central vision appears to be more efficient (on a per-pixel basis) than peripheral vision in all experiments we tried, we tested the relative efficiency of the central vision over peripheral vision by measuring the recognition accuracy as a function of viewable area. The result is shown in Fig- ure 7. From Figure 7, we can clearly see that the recognition ac- curacy of central vision is always superior than peripheral vision for all tasks. However, central vision is even more efficient when recognizing faces than recognizing objects or scenes, as viewable areas over 50% of the whole image can only provide a limited boost for face recognition, while sig- nificantly improving the accuracy of object and scene recog- nition. Contrarywise, peripheral information provides little to no help for face recognition, unless over 90% of the image is Figure 6: Results for large-scale face recognition accuracy as a function of viewing condition (Windows (w) and Sco- tomas (s)) and visual angle. Left: experiment with origi- nal images. Right: experiment with foveated images. For Lighten-A and Lighten-B models, the visual angle only ex- pands to 9.5◦, as the input image is smaller (144× 144) than for the VGG model (256×256). The accuracy for face verifi- cation task is measured as the true positive rate at Equal Error Rate (EER) point on the ROC curve. Chance is 0.5. presented, but the accuracy still suffers due to the loss of cen- tral vision. However, peripheral information is important for object and scene recognition (and more important for scene recognition, as shown in Figure 5). These large-scale scene, object, and face recognition mod- eling results suggest there is an order of relative importance of central versus peripheral vision in those tasks: peripheral vision is most important for scene recognition, less impor- tant for object recognition, and basically not helpful for face recognition. Central vision, however, plays a crucial role in face recognition, is important for object recognition, and is less important for scene recognition. Conclusion In this paper, we modeled the contribution of central ver- sus peripheral visual information for scene, object, and face recognition, using deep CNNs. We first modeled the behav- ioral study of Larson and Loschky (2009), and replicated their findings of the importance of peripheral vision in scene recog- nition. In addition, by running a large-scale scene, object, and face recognition simulation, our models make testable predic- tions for the relative order of importance for central versus peripheral vision for those tasks. Acknowledgments This work was supported by NSF grants IIS-1219252 and SMA 1041755 to GWC. PW was supported by a fellowship from Hewlett-Packard. References Epstein, R., Harris, A., Stanley, D., & Kanwisher, N. (1999). The parahippocampal place area: Recognition, navigation, or encoding? Neuron, 23(1), 115–125. of central versus peripheral vision to scene gist recognition. Journal of Vision, 9(10), 6. Loschky, L., McConkie, G., Yang, J., & Miller, M. (2005). The limits of visual resolution in natural scene viewing. Vi- sual Cognition, 12(6), 1057–1092. Malach, R., Levy, I., & Hasson, U. (2002). The topogra- phy of high-order human object areas. Trends in cognitive sciences, 6(4), 176–184. McCandliss, B. D., Cohen, L., & Dehaene, S. (2003). The visual word form area: expertise for reading in the fusiform gyrus. Trends in cognitive sciences, 7(7), 293–299. Nelson, W. W., & Loftus, G. R. (1980). The functional vi- sual field during picture viewing. Journal of Experimental Psychology: Human Learning and Memory, 6(4), 391. Parkhi, O. M., Vedaldi, A., & Zisserman, A. (2015). Deep face recognition. In British machine vision conference. Polyak, S. L. (1941). The retina. Ren, S., He, K., Girshick, R., & Sun, J. (2015). Faster r- cnn: Towards real-time object detection with region pro- In Advances in neural information pro- posal networks. cessing systems (pp. 91–99). Russakovsky, O., Deng, J., Su, H., Krause, J., Satheesh, S., Ma, S., . . . Fei-Fei, L. ImageNet Large Scale Visual Recognition Challenge. International Journal of Computer Vision (IJCV), 1-42. doi: 10.1007/s11263- 015-0816-y (2015, April). Simonyan, K., & Zisserman, A. (2014). Very deep convo- lutional networks for large-scale image recognition. arXiv preprint arXiv:1409.1556. Szegedy, C., Liu, W., Jia, Y., Sermanet, P., Reed, S., (2015, June). Going Anguelov, D., . . . Rabinovich, A. deeper with convolutions.. van Diepen, P. M., Wampers, M., & dYdewalle, G. (1998). Functional division of the visual field: Moving masks and moving windows. Eye guidance in reading and scene per- ception, 337–355. Wandell, B. A. (1995). Foundations of vision. Sinauer Asso- ciates. Wang, P., Malave, V., & Cipollini, B. (2015). Encoding voxels with deep learning. The Journal of Neuroscience, 35(48), 15769–15771. Wu, X., He, R., & Sun, Z. (2015). A lightened cnn for deep face representation. arXiv preprint arXiv:1511.02683. Yamins, D. L., Hong, H., Cadieu, C. F., Solomon, E. A., Seib- ert, D., & DiCarlo, J. J. (2014). Performance-optimized hi- erarchical models predict neural responses in higher visual cortex. Proceedings of the National Academy of Sciences, 111(23), 8619–8624. Zeiler, M. D., & Fergus, R. (2014). Visualizing and under- standing convolutional networks. In Computer vision–eccv 2014 (pp. 818–833). Springer. Zhou, B., Lapedriza, A., Xiao, J., Torralba, A., & Oliva, A. (2014). Learning deep features for scene recognition using In Advances in neural information pro- places database. cessing systems (pp. 487–495). Figure 7: Accuracy for object (left), scene (middle) and face (right) recognition as a function of the percentage of view- able area presented under Window (blue) and Scotoma (red) conditions, using original (solid line) and foveated images (dashed line). Gomez, J., Pestilli, F., Witthoft, N., Golarai, G., Liberman, A., Poltoratski, S., . . . Grill-Spector, K. (2015). Func- tionally defined white matter reveals segregated pathways in human ventral temporal cortex associated with category- specific processing. Neuron, 85(1), 216–227. Grill-Spector, K., & Malach, R. (2004). The human visual cortex. Annu. Rev. Neurosci., 27, 649–677. Guc¸lu, U., & van Gerven, M. A. (2015). Deep neural net- works reveal a gradient in the complexity of neural repre- sentations across the ventral stream. The Journal of Neuro- science, 35(27), 10005–10014. He, K., Zhang, X., Ren, S., & Sun, J. (2015). Deep residual learning for image recognition. arXiv preprint arXiv:1512.03385. Henderson, J. M., & Hollingworth, A. (1999). The role of fix- ation position in detecting scene changes across saccades. Psychological Science, 10(5), 438–443. Holmes, D. L., Cohen, K. M., Haith, M. M., & Morrison, F. J. (1977). Peripheral visual processing. Perception & Psychophysics, 22(6), 571–577. Huang, G. B., Ramesh, M., Berg, T., & Learned-Miller, E. (2007). Labeled faces in the wild: A database for study- ing face recognition in unconstrained environments (Tech. Rep. No. 07-49). University of Massachusetts, Amherst. Jia, Y., Shelhamer, E., Donahue, J., Karayev, S., Long, J., Girshick, R., . . . Darrell, T. (2014). Caffe: Convolutional architecture for fast feature embedding. arXiv preprint arXiv:1408.5093. Kanwisher, N., McDermott, J., & Chun, M. M. (1997). The fusiform face area: a module in human extrastriate cor- tex specialized for face perception. The Journal of Neu- roscience, 17(11), 4302–4311. Krizhevsky, A., Sutskever, I., & Hinton, G. E. (2012). Im- agenet classification with deep convolutional neural net- works. In Advances in neural information processing sys- tems (pp. 1097–1105). Larson, A. M., & Loschky, L. C. (2009). The contributions
1702.08617
1
1702
2017-02-28T03:03:40
Spontaneous Activity in the Visual Cortex is Organized by Visual Streams
[ "q-bio.NC" ]
Large-scale functional networks have been extensively studied using resting state functional magnetic resonance imaging. However, the pattern, organization, and function of fine-scale network activity remain largely unknown. Here we characterized the spontaneously emerging visual cortical activity by applying independent component analysis to resting state fMRI signals exclusively within the visual cortex. In this sub-system scale, we observed about 50 spatially independent components that were reproducible within and across subjects, and analyzed their spatial patterns and temporal relationships to reveal the intrinsic parcellation and organization of the visual cortex. We found that the visual cortical parcels were aligned with the steepest gradient of cortical myelination, and organized into functional modules segregated along the dorsal/ventral pathways and foveal/peripheral early visual areas. In contrast, cortical retinotopy, folding, and cytoarchitecture impose limited constraints to the organization of resting state activity. From these findings, we conclude that spontaneous activity patterns in the visual cortex are primarily organized by visual streams, likely reflecting feedback network interactions.
q-bio.NC
q-bio
Spontaneous Activity in the Visual Cortex is Organized by Visual Streams Kun-Han Lu2,3, Jun Young Jeong2, Haiguang Wen2,3, Zhongming Liu*1,2,3 1Weldon School of Biomedical Engineering 2School of Electrical and Computer Engineering 3Purdue Institute for Integrative Neuroscience Purdue University, West Lafayette, IN, USA *Correspondence Zhongming Liu, PhD Assistant Professor of Biomedical Engineering Assistant Professor of Electrical and Computer Engineering College of Engineering, Purdue University 206 S. Martin Jischke Dr. West Lafayette, IN 47907, USA Phone: +1 765 496 1872 Fax: +1 765 496 1459 Email: [email protected] 1 Abstract Large-scale functional networks have been extensively studied using resting state functional magnetic resonance imaging. However, the pattern, organization, and function of fine-scale network activity remain largely unknown. Here we characterized the spontaneously emerging visual cortical activity by applying independent component analysis to resting state fMRI signals exclusively within the visual cortex. In this sub-system scale, we observed about 50 spatially independent components that were reproducible within and across subjects, and analyzed their spatial patterns and temporal relationships to reveal the intrinsic parcellation and organization of the visual cortex. We found that the visual cortical parcels were aligned with the steepest gradient of cortical myelination, and organized into functional modules segregated along the dorsal/ventral pathways and foveal/peripheral early visual areas. In contrast, cortical retinotopy, folding, and cytoarchitecture impose limited constraints to the organization of resting state activity. From these findings, we conclude that spontaneous activity patterns in the visual cortex are primarily organized by visual streams, likely reflecting feedback network interactions. Keywords: Fine-scale networks; independent component analysis; functional parcellation; visual streams. 2 Introduction Resting state functional magnetic resonance imaging (rs-fMRI) has been widely explored to map resting state networks (RSNs) that collectively report the brain's intrinsic functional organization (Fox and Raichle, 2007). These networks consist of temporally correlated regions (Biswal et al., 1995; Van Dijk et al., 2010), arise from structural connections (Honey et al., 2007; Van Den Heuvel et al., 2009), resemble and predict task activations (Smith et al., 2009; Cole et al., 2014; Tavor et al., 2016), and distinguish individual subjects (Finn et al., 2015) or diseases (Fox and Greicius, 2010). Although brain activity spans a variety of spatial scales (Yoshimura et al., 2005; Doucet et al., 2011; Hutchison et al., 2013), RSNs have been mostly characterized in the whole brain, where their functions are empirically and coarsely defined in terms of motor (Biswal et al., 1995), vision (Yeo et al., 2011), default-mode (Greicius et al., 2003), attention (Fox et al., 2006), salience (Seeley et al., 2007), and so on. In finer spatial scales, patterns of spontaneous activity and connectivity remain unclear, but may bear more specific functional roles related to perception, behavior, or cognition (Kenet et al., 2003; Yoshimura et al., 2005; Wang et al., 2013; Long et al., 2014; Wilf et al., 2015; Lewis et al., 2016). In this regard, the visual cortex is a rich and ideal benchmark. It has been characterized in terms of cellular architectonics (Amunts et al., 2000; Glasser and Van Essen, 2011), structural connections (Felleman and Van Essen, 1991; Salin and Bullier, 1995), cortical folding (Fischl et al., 2008; Benson et al., 2012), functional pathways (Hubel and Wiesel, 1962; Ungerleider and Haxby, 1994), and neural coding (Ohiorhenuan et al., 2010; Guclu and van Gerven, 2015). In recent studies, correlation patterns of spontaneous activity in the visual cortex have been examined and compared with retinotopy (Heinzle et al., 2011; Yeo et al., 2011; Jo et al., 2012; Butt et al., 2013; de Zwart et al., 2013; Gravel et al., 2014; Raemaekers et al., 2014; Arcaro et al., 2015; Bock et al., 2015; Striem-Amit et al., 2015; Wilf et al., 2015; Dawson et al., 2016; Genc et al., 2016; Lewis et al., 2016). Since much of the visual cortex has visual- field maps – topographic representations of the polar angle and eccentricity (Wandell et al., 2007), it is reasonable to initially hypothesize that resting state activity is also intrinsically organized by retinotopy. 3 Support for this hypothesis comes in part from the finding that in early visual areas (e.g. V1-V3), correlations in spontaneous activity are generally higher between locations with similar eccentricity representations (Heinzle et al., 2011; Yeo et al., 2011; Jo et al., 2012; Gravel et al., 2014; Arcaro et al., 2015; Striem-Amit et al., 2015; Dawson et al., 2016; Genc et al., 2016; Lewis et al., 2016). However, such eccentricity-dependent functional connectivity has rarely been reported beyond early visual areas (Baldassano et al., 2013; Striem-Amit et al., 2015); it may be relatively weak (Wilf et al., 2015) or observable only after regressing out large-scale activity (Raemaekers et al., 2014); and it may also be confounded by the decay of local connectivity over cortical distance (Butt et al., 2013; Dawson et al., 2016). Further against this hypothesis, spontaneous activity preserves little or less polar-angle dependence in intra-hemispheric correlations (Gravel et al., 2014; Bock et al., 2015; Wilf et al., 2015), and shows strong correlations between bilateral V1 locations despite their lack of common receptive fields or direct connections (Jo et al., 2012; de Zwart et al., 2013). As such, the retinotopic organization of fine-scale resting activity in the visual cortex is questionable (Butt et al., 2013; Wilf et al., 2015). The previously reported eccentricity-dependent spontaneous activity and connectivity may be attributed to alternative representations that partly overlap with the eccentricity representation in early visual areas, as opposed to retinotopy per se. In fact, cortical representations of the peripheral and central visual fields partly overlap with the magnocellular and parvocellular streams (Schiller et al., 1990; Nassi and Callaway, 2009), and extend onto the dorsal and ventral pathways for visual action and perception, respectively (Goodale and Milner, 1992; Ungerleider and Haxby, 1994). Along these pathways, feedforward neuronal circuits convey and integrate not only visual positions, but also increasingly complex visual or conceptual features (Martin, 2007; Hasson et al., 2008; Yamins and DiCarlo, 2016). Top-down feedback connections are not or less retinotopically organized than are feedforward pathways (Salin and Bullier, 1995). The complex interplay between feedforward and feedback processes is essential for natural vision (Rao and Ballard, 1999; Gilbert and Li, 2013), but remains largely unclear in a stimulus-free resting state. Thus, spontaneously 4 emerging networks in the visual cortex may not readily fit the retinotopic organization, or arguably any other presumable organizations. What is needed is a data-driven analysis of resting-state activity in a finer spatial scale, unbiased by any presumed areal definition or organizational hypothesis. For this purpose, independent component analysis (ICA) is well suited but has not been applied to finer spatial scales, to our knowledge, despite its wide application to whole-brain fMRI signals (Damoiseaux et al., 2006). Unlike the correlation analysis (Baldassano et al., 2012; Genc et al., 2016), ICA also has the advantages of being multivariate and data- driven, thereby bypassing the potential bias from any narrowly-focused hypothesis (Calhoun et al., 2009). Here, we explored a new application of ICA for mapping cortical visual areas and networks based on rs- fMRI signals within the visual cortex. In this sub-system scale, the fine-grained activity patterns derived from ICA were systematically characterized, interpreted, and evaluated for their test-retest reproducibility and individual variations. Moreover, they were also compared against cortical folding (Destrieux et al., 2010), retinotopy (Abdollahi et al., 2014), cytoarchitecture (Van Essen et al., 2012a), myeloarchitecture (Glasser et al., 2014), and the latest multimodal cortical parcellation (Glasser et al., 2016). Clustering analysis further reveals that spontaneously emerging network patterns in the human visual cortex are not retinotopically organized; instead, they are temporally clustered into three functional modules: the dorsal pathway, the ventral pathway, and the foveal and peripheral sub-divisions of early visual areas. Materials and Methods Subjects and Data We used the rs-fMRI data released from the Human Connectome Project (HCP) (Van Essen et al., 2013). Briefly, we randomly selected 201 independent healthy subjects; for each subject, we used the 5 data from two resting state sessions (session 1 for test and session 2 for retest); each session was 14 minutes and 33 seconds with the eyes open and fixated. As elaborated elsewhere (Van Essen et al., 2012b), data were acquired in a 3-tesla MRI system with a 32-channel head coil (Skyra, Siemens, Germany). The rs-fMRI data we used were acquired with a single-shot, multiband-accelerated, gradient-recalled echo-planar imaging with nominally 2mm isotropic spatial resolution and 0.72s temporal resolution, and left-to-right phase-encoding. In addition, structural images with T1 and T2-weighted contrast were both acquired with 0.7mm isotropic resolution. As elaborated elsewhere (Glasser et al., 2013), the structural images were non-linearly registered to the Montreal Neurological Institute (MNI) template, where the images were combined and segmented to generate cortical surfaces. The fMRI images were corrected for slice timing and motion, aligned to structural images, normalized to the MNI space, projected onto the cortical surfaces, and co-registered across subjects. In addition to the minimal preprocessing described above, we removed the slow trend in the fMRI time series by regressing out a fourth-order polynomial function, and subtracted the mean and standardized the signal variance. Note that spatial smoothing was not performed to minimize spurious correlations in neighboring voxels. Independent Component Analysis We applied the ICA to the rs-fMRI signals within a mask of the visual cortex (Fig. 1.A), defined by a system-level functional parcellation of the human cortex (Yeo et al., 2011). The fMRI time series within the mask was temporally standardized and concatenated across subjects. Infomax ICA (Bell and Sejnowski, 1995) was used to decompose the concatenated rs-fMRI data into 59 spatially independent components within the visual cortex. Here, the number of independent components, 59, was determined by maximizing the Laplace approximation of the posterior probability of the ICA model order (Beckmann and Smith, 2004). The spatial pattern of each component was converted to a z-score map by dividing the IC weight at each voxel by the standard deviation of voxel-wise residual noise that could not be explained 6 by the ICA model (Beckmann and Smith, 2004). All color scale for displaying the spatial IC maps in this study are z-scores. Reproducibility test Furthermore, we evaluated the test-retest reproducibility of the ICA results. In a group level, the ICA applied to the data in one session (i.e. session 1) was also applied to the data in a repeated session (i.e. session 2); both sessions were from the same group of subjects, and the data was concatenated in the same order across subjects. We calculated the absolute value of the spatial correlation (r) between every component from session 1 and every component from session 2, and paired these components across sessions into distinct pairs to maximize the sum of their absolute spatial correlations. More specifically, we used an iterative procedure toward the optimal pairing: we began with identifying a pair of ICs (i.e. one from session1, and the other from session2) with the highest correlation; then we paired these two ICs, and excluded them from subsequent pairing, which continued until all ICs were paired. Note that the absolute spatial correlation was used since a reproducible IC could show the same spatial distribution despite opposite polarity. A threshold (r ≥ 0.4) was used to identify reproducible components for further interpretation (Fig. 1.C). We also explored the potential confounding effects of head motion on fine-scale independent components. For each subject, we regressed out the time series of six motion correction parameters prior to group-level ICA. We compared the ICs obtained without and with the above motion correction, and further identified and excluded those ICs that appeared inconsistent solely due to this preprocessing step. Modularity analysis For all the ICs that were reproducible and unaffected by head motion correction, we computed the temporal correlations between different components. Such correlations were first calculated based on component time series from each subject, and then averaged across subjects. This procedure prevented the 7 resulting correlations from being dominated or biased by inter-subject variations. To evaluate the statistical significance of the between-component correlation, we converted the correlation coefficient to the z-score using the Fisher's r-to-z transform separately for each subject, and then applied a one-sample t-test (with dof=200, significance level at 0.01, and Bonferroni correction) to the z-scores from all the subjects. We also applied the Louvain modularity analysis (Blondel et al., 2008) to the cross-component correlation matrix averaged across subjects. It assigned individual ICs to different modules, such that the temporal correlations were higher within modules but lower between modules. The modularity analysis, including the determination of the number of modules, was conducted with the algorithm (a Matlab function: modularity_louvain_und_sign) implemented in Brain Connectivity Toolbox (Rubinov and Sporns, 2010). The modularity index, Q, which quantified the goodness of modularity partitions (ranged from 0 to 1), was obtained. To evaluate the statistical significance of Q, we randomly shuffled the values within the correlation matrix 10,000 times and computed Q for each permutation given the same module assignment. This generated a null distribution of Q, against which the p value was computed for the Q value without permutation. To visualize the modular organization on the cortical surface, we represented each IC by a sphere located at the peak location in the component map, and color-coded every IC by its module membership. Computation of cortical distance To examine the effect of cortical distance on temporal correlations between components, we first identified the peak location in each component map (hereafter referred to as the component centroid). For bilaterally-distributed components, two centroids were identified: one at each hemisphere. We computed the geodesic cortical distances along the cortical surface between different component centroids by using HCP's "-surface-geodesic-distance" function. The cortical distances were computed for each subject and for each hemisphere, and then averaged across subjects. 8 The relationship between the temporal correlations and the corresponding cortical distances was examined and fitted by a rational function based on least-squares estimations (the "lsqcurvefit" function in Matlab). Separately for each hemisphere, we defined the spatial affinity between components as the reciprocal of the cortical distances between the corresponding centroids. Then we applied the Louvain modularity analysis to the spatial-affinity matrix in the same way as for the modularity analysis of functional connectivity. We further compared such anatomically-defined modules with the functional modules obtained on the basis of the temporal correlations between components. Intrinsic functional parcellation of the visual cortex From all the ICs, we created a group-level intrinsic functional parcellation of the visual cortex with an increasing level of granularity. Specifically, we defined a feature vector for each voxel in the visual cortex. This feature entails the weights by which the individual time series of different ICs were linearly combined to explain the fMRI signal observed at each voxel. Then we grouped the cortical voxels into distinct parcels by applying the k-means clustering to the corresponding feature vectors using a correlation-based "distance" and 1,000 replications with random initialization. The number of clusters (k) was empirically set to 10, 20, 30, 40. The parcels were color-coded from 0 to 1 in an ascending order according to the averaged "distance" within each parcel. We preferred this k-means clustering analysis to a "winner-take-all" alternative, in which each voxel was assigned to only one IC with the greatest weight (among all the ICs) at the given voxel. This was because single voxel time series were not necessarily represented by only one IC, but instead often by a few ICs. To facilitate interpretation, the ICA components and the parcellation derived from them were compared against conventional visual areas or cortical parcellation based on various structural and/or functional properties, including myeloarchitecture (Glasser et al., 2014), cytoarchitecture (Eickhoff et al., 2005), cortical folding (Destrieux et al., 2010), retinotopic mapping (Abdollahi et al., 2014), and multimodal parcellation (Glasser et al., 2016). 9 Dual-regression and individual-level parcellation Following group ICA, we also used dual regression (Filippini et al. 2009) against each subject's fMRI data to characterize subject-specific ICA maps (Tavor et al. 2016). Briefly, we first applied multiple regression to the spatial domain, using the group-level ICA spatial maps as a set of spatial regressors to obtain individual time series that was associated with each group-level spatial map based on the subject- specific fMRI data; after normalizing these individual-level time series to a zero mean and a unitary variance, we applied multiple regression to the time domain, by using the normalized individual time series as temporal regressors to obtain the subject-specific ICA spatial maps. From the individual-level ICA maps, we also used the k-means clustering analysis as mentioned above to create subject-specific parcellation of the visual cortex, and compared it against the group-level parcellation. Results Intrinsic activity patterns within the visual cortex Here, we explored a data-driven analysis of spontaneous activity confined to the human visual cortex. In this finer sub-system scale, our goal was to characterize and map intrinsic activity patterns in order to delineate cortical visual areas and networks independent of any task context or any presumed organizational hypothesis. Towards this goal, we used a cortical mask (Fig. 1.A), based on a systems- level parcellation of the entire cortex (Yeo et al., 2011), to only select rs-fMRI activity within the human visual cortex for group ICA. The selected rs-fMRI data from 201 healthy human subjects in the Human Connectome Project (HCP) were temporally standardized within each subject and concatenated across subjects. The concatenated data were then decomposed into 59 spatially independent components (ICs). Repeating this analysis with data from a different resting-state session for the same subjects allowed us to evaluate the test-retest reproducibility of every component. For example, Fig. 1.B displays three typical ICs that were spatially consistent (or correlated) between the two repeated sessions. By pairing the ICs across sessions to maximize the sum of the absolute pairwise correlation coefficients, we identified 50 10 unique pairs of reproducible components, which showed higher spatial correlations within pairs (r>0.4, mean±S.D = 0.78±0.14) than across pairs (mean±S.D = 0.04±0.05) (Fig. 1.C). In addition, we found that IC51 and IC55 yielded high spatial correlations with IC36 (r = 0.578) and IC41 (r = 0.63), respectively. But they were not paired, because IC36 and IC41 were better paired with other components while the pairing algorithm did not allow any duplication. We included IC51 and IC55 as reproducible components in subsequent analyses. These results suggest that the ICA-derived spontaneous activity patterns are robust and reproducible in a sub-system spatial scale. Fig. 2 shows the spatial maps of all 52 reproducible components in a descending order of their test-retest reproducibility. Among them, 92% showed focal patterns, and fewer ICs were distributed (IC#: 34, 35, 49, 50); 30% showed bilateral distributions (IC#: 4, 8, 17, 19, 23, 28, 30, 31, 33, 34, 35, 36, 46, 49, 50); 36% showed clearly anti-correlated patterns with well localized positivity and negativity (IC#: 2, 3, 5, 6, 7, 8, 13, 15, 17, 21, 27, 29, 30, 31, 34, 42, 49, 50). Some of these components not only aligned with existing anatomical borders (e.g. IC36 and IC50 aligned with V1/V2 border), but also aligned with regions with known functional properties (e.g. IC41 and IC55 both matched well with the region MT). Next, we attempted to identify the components that might be susceptible to artifacts related to head motion. We compared the ICs obtained without and with motion correction (i.e. regressing out head motion correction parameters from voxel time series). Among all the components shown in Fig. 2, three ICs (IC34, IC49, IC50) with distributed patterns did not match to any of the ICs (r < 0.15) obtained after motion correction (Supplementary Fig. 1A). Thus we attributed these ICs to head motion, and further excluded them for subsequent analyses. All other 49 ICs in Fig. 2 were one-to-one matched to the ICs after head motion correction, showing high spatial correlations for all matched pairs (Supplementary Fig. 1.B). In the following sections, we further segregated and interpreted these 49 ICs by comparing them to existing visual areas or networks. Comparing discrete ICA components with existing visual areas 11 Since the ICA-derived activity patterns mostly showed discrete regions with well-defined borders (Fig. 2), we further compared such discrete ICs with the visual areas defined with a recently published multi-modal parcellation (MMP) (Glasser et al., 2016). For the primary visual area (V1), four components were found to be sharply confined to V1 (Fig. 3.A). IC8 matched the bilateral foveal representations in V1; IC36 also showed bilateral distributions and corresponded to more peripheral representations; IC26 and IC38 showed unilateral distributions, corresponding to the most peripheral part of the right and left visual fields, respectively; these components did not overlap each other and all aligned with the V1 border. Therefore, V1 consists of multiple intrinsic functional sub-divisions apparently organized according to eccentricity representations, being largely symmetric not only between the left and right hemispheres, but also between the upper and lower sides of the calcarine sulcus. Unlike V1, V2 or V3 did not confine any component within itself. Instead, multiple components spanned across V2 and V3 along either the dorsal (IC18, IC30) or ventral (IC15 and IC9) direction (Fig. 3.B). Compared to those components within V1, the V2/V3 components were less bilaterally symmetric; none of them included regions in both dorsal and ventral pathways. Beyond those in early visual areas (V1/V2/V3), other discrete components were all anatomically split by the dorsal-ventral division. In the dorsal pathway, some components matched well with existing visual areas (Fig. 4A), including the middle temporal (MT: IC41 on the right hemisphere, IC55 one the left hemisphere), caudal area of inferior parietal cortex (PGp: IC19), dorsal visual transitional area (DVT: IC51 on the right hemisphere, IC39 on the left hemisphere) and parieto-occipital sulcus area 2(POS2: IC4). Some other components were distributed across multiple visual areas, including the third visual area and the area intraparietal 0 (V3B/IP0: IC44 on the right hemisphere, IC20 on the left hemisphere), third visual areas and the fourth visual area (V3A/V3B/V3CD/V4: IC31), the dorsal visual transitional area and the sixth visual area (DVT/V6A: IC28), and the third visual area, the sixth visual area and the seventh visual area (V3A/V6A/V7: IC23); about half of these dorsal components were bilaterally symmetric. Along the ventral pathway, the majority of the components covered multiple visual areas (Fig. 4B), including ventral-medial visual areas (VMV1/VMV2/VMV3: IC42 on the right hemisphere, IC22 on the 12 left hemisphere), para-hippocampal areas (PHA2/PHA3: IC10 on the right hemisphere, IC16 on the left hemisphere), fourth visual areas and the area lateral occipital 2 (V4t/LO2: IC33), the ventral visual complex and the fusiform face complex (VVC/FFC: IC40) and VMV2/PHA2/PHA3 (IC37), except for VMV1 (IC45); about two third of these ventral components were bilaterally symmetric; the rest of them were unilateral. Functional modularity in the visual cortex For all discrete components, we further evaluated their modular organization based on their temporal correlations. The between-component correlation matrix was calculated for every subject, and then averaged across subjects. The resulting group-level correlation matrix was re-organized into four functional modules based on the Louvain modularity analysis (Blondel et al., 2008; Rubinov and Sporns, 2010). Components within the same module were strongly and positively correlated, whereas components from different modules were weakly or negatively correlated (Fig. 5B, left). Visualizing the distributions of these functional modules on the cortical surface revealed their anatomical segregation (Fig. 5A). The first module included components, each of which was represented by one or two spheres for unilateral (or bilateral) components, over the foveal representations of early visual areas (V1, V2, and V3), whereas the second module was mostly distributed over the peripheral representations of early visual areas. The third module was distributed along the dorsal pathway, and the fourth module was distributed along the ventral pathway (Fig. 5.A). The modularity (Q=0.4966) was statistically significant (p < 0.0001, non-parametric permutation test). The between-component correlations within every module were consistently high for most of the subjects, yielding high t statistics, especially for the module over the foveal early visual areas and the module over the ventral visual areas. The t statistics corresponding to the correlations across different modules were generally low and not significant. These results lend support to the notion that intrinsic networks within the visual cortex are organized into functional modules: the dorsal pathway, the ventral pathway, as well as the foveal and peripheral parts of the early visual areas. 13 Effects of cortical distance on functional connectivity and modularity The observation that the functional modules were anatomically clustered led us to ask whether the functional relationships between individual ICs were entirely attributable to their cortical distances. To address this question, we tried to model the temporal correlations between components as a function of the cortical-surface distance between component centroids, separately evaluated for each hemisphere. The scatter-plots of correlation (r) vs. distance (d) revealed reciprocal relationships, which were modeled as r=1.432/d and r=1.521/d for left and right hemispheres, respectively, based on least-squares estimation (Fig. 6, left). It suggests that fine-scale functional connectivity within a cortical hemisphere depends, at least in part, on cortical distance, in line with findings from previous studies (Dawson et al., 2016; Genc et al., 2016). Despite this relationship, we further asked whether the modular organization based on functional connectivity (Fig. 5) could be a result of spatial affinity, described as the reciprocal of the distance along the cortical surface. The matrix of spatial affinity was computed for each hemisphere, and reorganized into modules using the same modularity analysis as used for functional connectivity (Fig. 6, right). The left hemisphere contained two anatomical modules: one covering the lateral cortical surface, and the other covering the medial surface. The right hemisphere exhibited similar anatomical modules as the left hemisphere, except that the medial cortical surface was sub-divided into two modules in the right hemisphere. The modularity (Q) of the spatial affinity matrices was 0.2039 and 0.2053, which were statistically significant (p<0.01) but notably lower than the modularity of the functional connectivity (Q=0.4966). More importantly, the functional and anatomical modules, on the basis of temporal correlations and spatial affinity respectively, showed different spatial distributions, as comparatively shown in Fig. 5 and Fig. 6. Thus, spontaneous functional connectivity in the visual cortex revealed a stronger and different modular organization and distribution than what were attainable based on spatial affinity alone. 14 Functional parcellation of the visual cortex Following ICA, we applied the k-means clustering to the ICA weighting vectors at individual locations within the visual cortex, yielding an automated intrinsic parcellation of the visual cortex with a varying level of granularity with the number of clusters (k) being 10, 20, 30, 40. We found that as the number of clusters increased, coarser parcels were progressively sub-divided into finer parcels (Fig. 7). We settled at k=40, which roughly matched the expected number of visual areas (Glasser et al., 2016), and generated a set of well-defined and bilaterally symmetric parcels (Fig. 8). This parcellation based on spontaneous activity was further compared against existing parcellations of the visual cortex, based on the whole-brain multimodal images (i.e. MMP) (Glasser et al., 2016), visual-field maps (Abdollahi et al., 2014), cortical folding patterns (Destrieux et al., 2010), cortical cytoarchitecture (Eickhoff et al., 2005), and cortical myelination (Glasser et al., 2014). It was found that none of the existing parcellations precisely agreed with the fully-automated and data-driven parcellation reported here. In particular, our parcellation did not match with those based on cortical retinotopy, cytoarchitecture, and folding, but matched relatively better with the cortical myelination and the MMP. In our parcellation, the outer contour of cortical parcels that covered early visual areas tended to align well with the steep gradients of cortical myelination. Our reported cortical parcels that covered the high-level visual areas tended to align reasonably well with the corresponding parcels in MMP; the alignment was not one to one. Compared to MMP, our parcellation was coarser in high-level visual areas, but finer in low-level visual areas. Parcellation in the level of single subjects In addition to the group-level analysis, we also applied dual regression to the dataset, in an attempt to obtain the corresponding ICA patterns from individual subjects. The individual-level ICA patterns were generally consistent with the group-level ICA patterns and comparable across subjects (Fig. 9.A). On the basis of individual-specific ICs, the visual cortex could be parcellated (k = 40) for each subject, yielding cortical parcels apparently noisy but generally similar to the group-level parcellation (Fig. 9.B). 15 Discussion We characterized the network patterns emerging from spontaneous resting-state activity within the human visual cortex. On the basis of such patterns and their interactions, we delineated the intrinsic functional parcellation and organization of the visual cortex. Here we report that fine-scale intrinsic visual cortical networks are not organized by the patterns of cortical retinotopy, folding, or cytoarchitecture, but align with the gradient of cortical myelination, and are segregated into functional modules specific to the ventral and dorsal visual streams. Whole-brain vs. fine-scale functional networks The majority of resting-state fMRI literature focuses on intrinsic functional networks in the whole brain scale (Smith et al., 2013). Large-scale networks are supported by long-range structural connections (Yeo et al., 2011), and reflect coarsely defined functions (Smith et al., 2009). However, in a finer scale within the visual cortex, bi-directional structural connections co-exist over short distances (Felleman and Van Essen, 1991; Salin and Bullier, 1995), forming the network basis of vision (Rao and Ballard, 1999). The patterns and dynamics of fine-scale intrinsic networks may indicate how visual information is coded in spontaneous brain activity (Kenet et al., 2003), and offer a more specific clue on the functional role of spontaneous activity in shaping perception or behavior (Wilf et al., 2015). To explore the topographic organization of fine-scale visual networks, it is necessary to confine the analysis to the functional connectivity profile within the visual system. Otherwise, in the whole-brain scale, the connectivity profile between a seed location in the visual cortex and the rest of the brain, as a function of locations, includes mostly the remote locations that are not or non-specifically associated with the seed location. As a result, the difference in the whole-brain connectivity profile of two distinct seed locations becomes subtle, even if they may actually interact with different sets of vison-related brain 16 locations. Focusing on a fine scale improves the sensitivity to differentiate the topographic difference in functional connectivity of specific interest to vision. In line with this notion, previous studies have shown that fine-scale functional networks may be obscured by large-scale network activity (Raemaekers et al., 2014), and thus appear to exhibit a coarse topographic organization (Nir et al., 2006; Yeo et al., 2011). Within the visual cortex, we used data-driven ICA to explore the multivariate voxel patterns and dynamics, instead of bivariate correlations between voxels or areas, as in previous studies (Heinzle et al., 2011; Yeo et al., 2011; Butt et al., 2013; de Zwart et al., 2013; Gravel et al., 2014; Raemaekers et al., 2014; Arcaro et al., 2015; Bock et al., 2015; Striem-Amit et al., 2015; Wilf et al., 2015; Dawson et al., 2016; Genc et al., 2016). In the whole-brain scale, ICA and seed-based correlation analyses have been shown to reveal similar network patterns (Van Dijk et al., 2010). However, structural connections are much denser at a reduced spatial scale or distance (Bassett and Bullmore, 2006), giving rise to more complex patterns of functional interactions. This makes ICA a more preferable method for fine-scale network mapping. Fine-scale visual networks are not retinotopically organized Our data suggest that resting state networks within the visual cortex are robust in group (Fig. 1) and individual (Fig. 9) levels, and well-organized in space (Figs. 2 through 4) and time (Figs. 5, 6). Here, the results obtained with data-driven ICA extend the previous findings obtained by correlational analyses of rs-fMRI signals in the visual cortex. However, unlike some prior studies (Heinzle et al., 2011; Gravel et al., 2014; Raemaekers et al., 2014), we did not find any evidence for the retinotopic organization of resting state activity beyond early visual areas. Perhaps, the exception was only in V1, where activity patterns were found to agree with eccentricity representations (Fig. 3A), consistent with previous findings (Yeo et al., 2011; Arcaro et al., 2015; Wilf et al., 2015). Beyond V1, spontaneous activity patterns were independent of either the eccentricity or the polar angle (Fig. 3). Even in V1, resting state activity was correlated between the left 17 and right hemispheres (Fig. 3), although the two hemispheres correspond to different hemi-fields in the visual space. Note that the left and right V1 areas have little or no callosal connections (Tootell et al., 1998) to directly support their synchronization. The inter-hemispherical V1 correlation is most likely due to a common input to both hemispheres. The topography of this common input seems retinotopically non- specific, at least in terms of the polar angle. Arguably, the eccentricity-dependent intrinsic activity patterns in V1 may be coincidental, and reflect the relative distributions of magnocellular (M) and parvocellular (P) projections that happen to vary with eccentricity. Previous studies have shown that P cells, relative to M cells, over-represent the central vision but under-represent the periphery in lateral geniculate nuclei (LGN) (Connolly and Van Essen, 1984; Schiller et al., 1990) and V1 (Baseler and Sutter, 1997; Azzopardi et al., 1999). The ratio between P and M projections to/from V1 notably decreases with eccentricity (Baseler and Sutter, 1997). Note that P and M pathways convey distinct visual attributes, but share the same retinotopic maps (Nassi and Callaway, 2009; Denison et al., 2014). Although the representation of the M-to-P ratio seems similar as the eccentricity representation in V1, the M-P pathways bear a different organization specific to visual streams as opposed to visual locations. Caution should be exercised before interpreting an eccentricity- dependent pattern alone as evidence for the retinotopic organization. Resting-state activity reflects feedback visual-network interactions We further speculate that the common input to V1, which drives visual-stream specific resting- state activity, arises from top-down modulations through feedback connections. In the visual hierarchy (Felleman and Van Essen, 1991), the population receptive field becomes larger and less specific from lower to higher visual areas (Wandell et al., 2007), to progressively converge information in the visual space through feedforward connections. While the feedforward connections are retinotopically organized, the feedback corticocortical connections are not so (Salin and Bullier, 1995). Through feedback, the top- down modulations transfer information about a large or even the whole visual field to cortical locations 18 with specific receptive fields (Salin and Bullier, 1995), driving network activity patterns away from being retinotopically specific. It is conceivable that the organization of feedback connections plays a defining role to resting state networks within the visual cortex. This is because when feedforward pathways are not driven by fluctuating external inputs, feedback pathways are modulated by the brain's intrinsic activity. In addition, feedback connections in the visual system are generally separated by the ventral and dorsal pathways (Salin and Bullier, 1995; Gilbert and Li, 2013), functionally specialized for recognition and action, respectively (Goodale and Milner, 1992; Ungerleider and Haxby, 1994). Such a visual-stream- specific organization also applies to feedback connections from V1 to LGN (Briggs and Usrey, 2009). It lends support for our interpretation that feedback projections serve the structural network basis of the observed modular organization of intrinsic fine-scale functional networks distributed along the ventral and dorsal pathways (Fig. 5). As the ventral and dorsal pathways become intricate in early visual areas, fine-scale network patterns in V1/V2/V3 constitute the third functional module (Fig. 5). Intrinsic functional parcellation of the visual cortex On the basis of fine-scale resting-state activity patterns, the functional parcellation of the visual cortex appeared to be notably different from other parcellations based on cortical folding, retinotopy, and cytoarchitecture. The discrepancy with the cortical folding is perhaps reasonable, because the relationship between cortical morphology and functional organization is elusive and indirect, despite a developmental linkage likely between them as proposed elsewhere (Benson et al., 2012; Ronan and Fletcher, 2015). The discrepancy with visual field maps is perhaps also understandable, for the reasons elaborated in previous sections. In addition, the parcellation based on functional connectivity is applicable to all locations in the visual cortex, whereas the visual field is not mapped at all visual areas. For the apparently different parcellations based on regional cellular composition and inter-regional functional connectivity suggests a lack of one-to-one relationships between cell types and functional networks. 19 Our results show that the functional networks and parcels seem to align with the gradient of myelination (Fig. 8). We speculate that spontaneous activity shapes myelin density. It has been shown that electrical activity may promote myelination(Gibson et al., 2014), and that functional organization and cortical myelination may co-vary given plasticity (Fields, 2015; Hunt et al., 2016). As such, greater myelin density may imply greater functional specificity but less plasticity: early sensory areas are more functionally specific with greater myelination or less plasticity, whereas higher-order or multi-sensory areas are less functionally specific with less myelin density and less plasticity (Glasser et al., 2014). Acknowledgments This work was supported in part by NIH R01MH104402 (Z Liu) and Purdue University. Authors have no conflict of interest. 20 Figure Caption 21 Figure 1. Visual cortical mask and reproducibility of fine-scale network patterns. A. The visual cortex mask (in blue) is illustrated on inflated and flattened cortical surfaces. Reference lines mark the occipital-parietal sulcus (red dash line), the calcarine sulcus (purple dash line), a rough ventral-dorsal division (yellow dash line), from where the ventral and dorsal pathways are along the green and red arrows. B. Three examples of reproducible ICs that exhibit high correlations (r) in their spatial patterns. The color scale represents the z-score and is ranged from -0.1 to 0.1. C. Shown in the left is the spatial correlation matrix between the ICs obtained from two repeated resting-state sessions. The diagonal elements correspond to uniquely paired ICs; the off-diagonal elements are between unpaired ICs. The red line represents a correlation threshold (r=0.4), by which a pair of ICs was considered reproducible between sessions. Shown in the middle is the p-values associated with the spatial correlations. Shown in the right panel is the discrete histograms of the correlations for the paired (red) and unpaired (blue) ICs. Figure 2. 52 reproducible independent components with spatial correlation greater than 0.4. A. The ICs were numbered in a descending order of their test-retest reproducibility. IC51 and IC55 were not 22 optimally matched by our algorithm, but they yielded spatial correlations of 0.58 and 0.63 with IC39 and IC41 respectively. Therefore, they are also considered to be reproducible. B. The remaining ICs that have less consistency across the two datasets. Figure 3. Discrete ICs within the early visual areas (V1/V2/V3). A. Four ICs within V1. B. Four ICs within V2 and V3. The green lines are the borders of the multi-modal parcellation (Glasser et al., 2016), where V1/V2/V3 are defined. 23 Figure 4. Discrete ICs along the dorsal pathway and ventral pathway. A. The left panel shows the ICs that match well with existing lateral visual areas in the dorsal pathway. The right panel shows the ICs that match well with existing medial visual areas in the dorsal pathway. B. Example ICs that match well with existing medial visual areas in the ventral pathway such as VMV1-3, FFC and VVC. The green lines mark the existing visual areal borders according to the MMP. 24 Figure 5. Functional modules in the visual cortex. A. Each component is represented with a sphere, colored coded by its modular membership, and placed at the peak location of its spatial map. B. The matrix on the left shows the temporal correlations (r) between ICs, organized into four modules (in (1) foveal early visual cortex, (2) peripheral visual cortex, (3) dorsal pathway, (4) ventral pathway). The matrix on the right shows the t-statistics associated with the temporal correlations (z) between ICs, organized in the same order as the matrix on the left. 25 Figure 6. The effect of cortical distance on functional connectivity and modularity. A. Left panel: The scatter-plots of correlation (r) vs. distance (d), which were modeled as r=1.432/d based on least- squares estimation, in the left hemisphere. Right panel: The matrix shows the spatial-affinity between component centroids within the left hemisphere. Middle panel: Each component is represented with a sphere, colored coded by its modular membership, and placed at the peak location of its spatial map. B. Left panel: The scatter-plots of correlation (r) vs. distance (d), which were modeled as r=1.521/d based on least-squares estimation, in the right hemisphere. Right panel: The matrix shows the spatial-affinity between component centroids within the right hemisphere. Middle panel: Each component is represented with a sphere, colored coded by its modular membership, and placed at the peak location of its spatial map. 26 Figure 7. Functional parcellation of the visual cortex with varying number of parcels (k=10, 20, 30, 40). The parcels are color-coded from 0 to 1, sorted in an ascending order according to the averaged sum of distances within each parcel. 27 Figure 8. Comparing our parcellation (k = 40) with existing visual cortex parcellations. Our functional parcellation is compared against (1) whole-brain multimodal images (i.e. MMP) (Glasser et al., 2016); (2) visual-field maps (Abdollahi et al., 2014); (3) visual-field maps (Abdollahi et al., 2014); (4) cortical folding patterns (Destrieux et al., 2010); (5) cortical myelination (Glasser et al., 2014). 28 Figure 9. Independent components and parcellations from three individual subjects obtained through dual regression. A. Five example ICs from three subjects. B. The individual-level parcellations obtained from three example subjects. Supplementary Figure 1. 54 independent components obtained from motion-controlled fMRI dataset with spatial correlation greater than 0.15. A. The 54 ICs are spatially consistent with those observed in Fig. 2. B. Spatial cross correlation between ICs with and without controlling for motion effects. 29 References Abdollahi RO, Kolster H, Glasser MF, Robinson EC, Coalson TS, Dierker D, Jenkinson M, Van Essen DC, Orban GA (2014) Correspondences between retinotopic areas and myelin maps in human visual cortex. Neuroimage 99:509-524. Amunts K, Malikovic A, Mohlberg H, Schormann T, Zilles K (2000) Brodmann's areas 17 and 18 brought into stereotaxic space-where and how variable? Neuroimage 11:66-84. Arcaro MJ, Honey CJ, Mruczek RE, Kastner S, Hasson U (2015) Widespread correlation patterns of fMRI signal across visual cortex reflect eccentricity organization. Elife 4. Azzopardi P, Jones KE, Cowey A (1999) Uneven mapping of magnocellular and parvocellular projections from the lateral geniculate nucleus to the striate cortex in the macaque monkey. Vision research 39:2179-2189. Baldassano C, Beck DM, Fei-Fei L (2013) Differential connectivity within the Parahippocampal Place Baldassano C, Iordan MC, Beck DM, Fei-Fei L (2012) Voxel-level functional connectivity using spatial Baseler H, Sutter E (1997) M and P components of the VEP and their visual field distribution. Vision Area. Neuroimage 75:228-237. regularization. Neuroimage 63:1099-1106. research 37:675-690. Bassett DS, Bullmore E (2006) Small-world brain networks. Neuroscientist 12:512-523. Beckmann CF, Smith SM (2004) Probabilistic independent component analysis for functional magnetic resonance imaging. IEEE Trans Med Imaging 23:137-152. Bell AJ, Sejnowski TJ (1995) An information-maximization approach to blind separation and blind deconvolution. Neural Comput 7:1129-1159. Benson NC, Butt OH, Datta R, Radoeva PD, Brainard DH, Aguirre GK (2012) The retinotopic organization of striate cortex is well predicted by surface topology. Current biology : CB 22:2081-2085. Biswal B, Zerrin Yetkin F, Haughton VM, Hyde JS (1995) Functional connectivity in the motor cortex of resting human brain using echo-planar mri. Magnetic resonance in medicine 34:537-541. Blondel VD, Guillaume JL, Lambiotte R, Lefebvre E (2008) Fast unfolding of communities in large networks. Journal of Statistical Mechanics-Theory and Experiment 2008:P10008. Bock AS, Binda P, Benson NC, Bridge H, Watkins KE, Fine I (2015) Resting-State Retinotopic Organization in the Absence of Retinal Input and Visual Experience. The Journal of neuroscience : the official journal of the Society for Neuroscience 35:12366-12382. Briggs F, Usrey WM (2009) Parallel processing in the corticogeniculate pathway of the macaque monkey. Neuron 62:135-146. Butt OH, Benson NC, Datta R, Aguirre GK (2013) The fine-scale functional correlation of striate cortex in sighted and blind people. The Journal of neuroscience : the official journal of the Society for Neuroscience 33:16209-16219. Calhoun VD, Liu J, Adali T (2009) A review of group ICA for fMRI data and ICA for joint inference of imaging, genetic, and ERP data. Neuroimage 45:S163-172. Cole MW, Bassett DS, Power JD, Braver TS, Petersen SE (2014) Intrinsic and task-evoked network architectures of the human brain. Neuron 83:238-251. Connolly M, Van Essen D (1984) The representation of the visual field in parvicellular and magnocellular layers of the lateral geniculate nucleus in the macaque monkey. J Comp Neurol 226:544-564. Damoiseaux JS, Rombouts SA, Barkhof F, Scheltens P, Stam CJ, Smith SM, Beckmann CF (2006) Consistent resting-state networks across healthy subjects. Proceedings of the National Academy of Sciences of the United States of America 103:13848-13853. Dawson DA, Lam J, Lewis LB, Carbonell F, Mendola JD, Shmuel A (2016) Partial Correlation-Based Retinotopically Organized Resting-State Functional Connectivity Within and Between Areas of the Visual Cortex Reflects More Than Cortical Distance. Brain connectivity 6:57-75. 30 de Zwart JA, van Gelderen P, Liu Z, Duyn JH (2013) Independent sources of spontaneous BOLD fluctuation along the visual pathway. Brain topography 26:525-537. Denison RN, Vu AT, Yacoub E, Feinberg DA, Silver MA (2014) Functional mapping of the magnocellular and parvocellular subdivisions of human LGN. Neuroimage 102 Pt 2:358-369. Destrieux C, Fischl B, Dale A, Halgren E (2010) Automatic parcellation of human cortical gyri and sulci using standard anatomical nomenclature. Neuroimage 53:1-15. Doucet G, Naveau M, Petit L, Delcroix N, Zago L, Crivello F, Jobard G, Tzourio-Mazoyer N, Mazoyer B, Mellet E, Joliot M (2011) Brain activity at rest: a multiscale hierarchical functional organization. Journal of neurophysiology 105:2753-2763. Eickhoff SB, Stephan KE, Mohlberg H, Grefkes C, Fink GR, Amunts K, Zilles K (2005) A new SPM toolbox for combining probabilistic cytoarchitectonic maps and functional imaging data. Neuroimage 25:1325-1335. Felleman DJ, Van Essen DC (1991) Distributed hierarchical processing in the primate cerebral cortex. Fields RD (2015) A new mechanism of nervous system plasticity: activity-dependent myelination. Nature Cerebral cortex 1:1-47. reviews Neuroscience 16:756-767. Finn ES, Shen X, Scheinost D, Rosenberg MD, Huang J, Chun MM, Papademetris X, Constable RT (2015) Functional connectome fingerprinting: identifying individuals using patterns of brain connectivity. Nature neuroscience 18:1664-1671. Fischl B, Rajendran N, Busa E, Augustinack J, Hinds O, Yeo BT, Mohlberg H, Amunts K, Zilles K (2008) Cortical folding patterns and predicting cytoarchitecture. Cerebral cortex 18:1973-1980. Fox MD, Raichle ME (2007) Spontaneous fluctuations in brain activity observed with functional magnetic resonance imaging. Nature reviews Neuroscience 8:700-711. Fox MD, Greicius M (2010) Clinical applications of resting state functional connectivity. Frontiers in systems neuroscience 4:19. Fox MD, Corbetta M, Snyder AZ, Vincent JL, Raichle ME (2006) Spontaneous neuronal activity distinguishes human dorsal and ventral attention systems. Proceedings of the National Academy of Sciences 103:10046-10051. Genc E, Scholvinck ML, Bergmann J, Singer W, Kohler A (2016) Functional Connectivity Patterns of Visual Cortex Reflect its Anatomical Organization. Cerebral cortex 26:3719-3731. Gibson EM, Purger D, Mount CW, Goldstein AK, Lin GL, Wood LS, Inema I, Miller SE, Bieri G, Zuchero JB, Barres BA, Woo PJ, Vogel H, Monje M (2014) Neuronal activity promotes oligodendrogenesis and adaptive myelination in the mammalian brain. Science 344:1252304. Gilbert CD, Li W (2013) Top-down influences on visual processing. Nature reviews Neuroscience 14:350-363. Glasser MF, Van Essen DC (2011) Mapping human cortical areas in vivo based on myelin content as revealed by T1- and T2-weighted MRI. The Journal of neuroscience : the official journal of the Society for Neuroscience 31:11597-11616. Glasser MF, Goyal MS, Preuss TM, Raichle ME, Van Essen DC (2014) Trends and properties of human cerebral cortex: correlations with cortical myelin content. Neuroimage 93 Pt 2:165-175. Glasser MF, Coalson TS, Robinson EC, Hacker CD, Harwell J, Yacoub E, Ugurbil K, Andersson J, Beckmann CF, Jenkinson M, Smith SM, Van Essen DC (2016) A multi-modal parcellation of human cerebral cortex. Nature 536:171-178. Glasser MF, Sotiropoulos SN, Wilson JA, Coalson TS, Fischl B, Andersson JL, Xu J, Jbabdi S, Webster M, Polimeni JR, Van Essen DC, Jenkinson M, Consortium WU-MH (2013) The minimal preprocessing pipelines for the Human Connectome Project. Neuroimage 80:105-124. Goodale MA, Milner AD (1992) Separate visual pathways for perception and action. Trends Neurosci 15:20-25. Gravel N, Harvey B, Nordhjem B, Haak KV, Dumoulin SO, Renken R, Curcic-Blake B, Cornelissen FW (2014) Cortical connective field estimates from resting state fMRI activity. Front Neurosci 8:339. 31 Greicius MD, Krasnow B, Reiss AL, Menon V (2003) Functional connectivity in the resting brain: a network analysis of the default mode hypothesis. Proceedings of the National Academy of Sciences 100:253-258. Guclu U, van Gerven MA (2015) Deep Neural Networks Reveal a Gradient in the Complexity of Neural Representations across the Ventral Stream. The Journal of neuroscience : the official journal of the Society for Neuroscience 35:10005-10014. Hasson U, Yang E, Vallines I, Heeger DJ, Rubin N (2008) A hierarchy of temporal receptive windows in human cortex. The Journal of neuroscience : the official journal of the Society for Neuroscience 28:2539-2550. Heinzle J, Kahnt T, Haynes JD (2011) Topographically specific functional connectivity between visual field maps in the human brain. Neuroimage 56:1426-1436. Honey CJ, Kotter R, Breakspear M, Sporns O (2007) Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences of the United States of America 104:10240-10245. Hubel DH, Wiesel TN (1962) Receptive fields, binocular interaction and functional architecture in the cat's visual cortex. J Physiol 160:106-154. Hunt BA, Tewarie PK, Mougin OE, Geades N, Jones DK, Singh KD, Morris PG, Gowland PA, Brookes MJ (2016) Relationships between cortical myeloarchitecture and electrophysiological networks. Proceedings of the National Academy of Sciences 113:13510-13515. Hutchison RM, Womelsdorf T, Allen EA, Bandettini PA, Calhoun VD, Corbetta M, Della Penna S, Duyn JH, Glover GH, Gonzalez-Castillo J (2013) Dynamic functional connectivity: promise, issues, and interpretations. Neuroimage 80:360-378. Jo HJ, Saad ZS, Gotts SJ, Martin A, Cox RW (2012) Quantifying agreement between anatomical and functional interhemispheric correspondences in the resting brain. PloS one 7:e48847. Kenet T, Bibitchkov D, Tsodyks M, Grinvald A, Arieli A (2003) Spontaneously emerging cortical representations of visual attributes. Nature 425:954-956. Lewis CM, Bosman CA, Womelsdorf T, Fries P (2016) Stimulus-induced visual cortical networks are recapitulated by spontaneous local and interareal synchronization. Proceedings of the National Academy of Sciences of the United States of America 113:E606-615. Long X, Goltz D, Margulies DS, Nierhaus T, Villringer A (2014) Functional connectivity-based parcellation of the human sensorimotor cortex. The European journal of neuroscience 39:1332- 1342. Martin A (2007) The representation of object concepts in the brain. Annu Rev Psychol 58:25-45. Nassi JJ, Callaway EM (2009) Parallel processing strategies of the primate visual system. Nature reviews Neuroscience 10:360-372. Nir Y, Hasson U, Levy I, Yeshurun Y, Malach R (2006) Widespread functional connectivity and fMRI fluctuations in human visual cortex in the absence of visual stimulation. Neuroimage 30:1313- 1324. Ohiorhenuan IE, Mechler F, Purpura KP, Schmid AM, Hu Q, Victor JD (2010) Sparse coding and high- order correlations in fine-scale cortical networks. Nature 466:617-621. Raemaekers M, Schellekens W, van Wezel RJ, Petridou N, Kristo G, Ramsey NF (2014) Patterns of resting state connectivity in human primary visual cortical areas: a 7T fMRI study. Neuroimage 84:911-921. Rao RPN, Ballard DH (1999) Predictive coding in the visual cortex: a functional interpretation of some extra-classical receptive-field effects. Nature neuroscience 2:79-87. Ronan L, Fletcher PC (2015) From genes to folds: a review of cortical gyrification theory. Brain structure Rubinov M, Sporns O (2010) Complex network measures of brain connectivity: uses and interpretations. & function 220:2475-2483. Neuroimage 52:1059-1069. Physiol Rev 75:107-154. Salin PA, Bullier J (1995) Corticocortical connections in the visual system: structure and function. 32 Schiller PH, Logothetis NK, Charles ER (1990) Functions of the colour-opponent and broad-band channels of the visual system. Nature 343:68-70. Seeley WW, Menon V, Schatzberg AF, Keller J, Glover GH, Kenna H, Reiss AL, Greicius MD (2007) Dissociable intrinsic connectivity networks for salience processing and executive control. The Journal of neuroscience 27:2349-2356. Smith SM, Fox PT, Miller KL, Glahn DC, Fox PM, Mackay CE, Filippini N, Watkins KE, Toro R, Laird AR, Beckmann CF (2009) Correspondence of the brain's functional architecture during activation and rest. Proceedings of the National Academy of Sciences of the United States of America 106:13040-13045. Smith SM, Vidaurre D, Beckmann CF, Glasser MF, Jenkinson M, Miller KL, Nichols TE, Robinson EC, Salimi-Khorshidi G, Woolrich MW, Barch DM, Ugurbil K, Van Essen DC (2013) Functional connectomics from resting-state fMRI. Trends in cognitive sciences 17:666-682. Striem-Amit E, Ovadia-Caro S, Caramazza A, Margulies DS, Villringer A, Amedi A (2015) Functional connectivity of visual cortex in the blind follows retinotopic organization principles. Brain : a journal of neurology 138:1679-1695. Tavor I, Parker Jones O, Mars RB, Smith SM, Behrens TE, Jbabdi S (2016) Task-free MRI predicts individual differences in brain activity during task performance. Science 352:216-220. Tootell RB, Mendola JD, Hadjikhani NK, Liu AK, Dale AM (1998) The representation of the ipsilateral visual field in human cerebral cortex. Proceedings of the National Academy of Sciences of the United States of America 95:818-824. Ungerleider LG, Haxby JV (1994) 'What' and 'where' in the human brain. Curr Opin Neurobiol 4:157- Van Den Heuvel MP, Mandl RC, Kahn RS, Pol H, Hilleke E (2009) Functionally linked resting-state networks reflect the underlying structural connectivity architecture of the human brain. Human brain mapping 30:3127-3141. Van Dijk KR, Hedden T, Venkataraman A, Evans KC, Lazar SW, Buckner RL (2010) Intrinsic functional connectivity as a tool for human connectomics: theory, properties, and optimization. Journal of neurophysiology 103:297-321. Van Essen DC, Glasser MF, Dierker DL, Harwell J, Coalson T (2012a) Parcellations and hemispheric asymmetries of human cerebral cortex analyzed on surface-based atlases. Cerebral cortex 22:2241-2262. Van Essen DC, Smith SM, Barch DM, Behrens TE, Yacoub E, Ugurbil K, Consortium WU-MH (2013) The WU-Minn Human Connectome Project: an overview. Neuroimage 80:62-79. Van Essen DC et al. (2012b) The Human Connectome Project: a data acquisition perspective. Neuroimage 62:2222-2231. 165. Wandell BA, Dumoulin SO, Brewer AA (2007) Visual field maps in human cortex. Neuron 56:366-383. Wang Z, Chen LM, Negyessy L, Friedman RM, Mishra A, Gore JC, Roe AW (2013) The relationship of anatomical and functional connectivity to resting-state connectivity in primate somatosensory cortex. Neuron 78:1116-1126. Wilf M, Strappini F, Golan T, Hahamy A, Harel M, Malach R (2015) Spontaneously Emerging Patterns in Human Visual Cortex Reflect Responses to Naturalistic Sensory Stimuli. Cerebral cortex. Yamins DL, DiCarlo JJ (2016) Using goal-driven deep learning models to understand sensory cortex. Nature neuroscience 19:356-365. Yeo BT, Krienen FM, Sepulcre J, Sabuncu MR, Lashkari D, Hollinshead M, Roffman JL, Smoller JW, Zollei L, Polimeni JR, Fischl B, Liu H, Buckner RL (2011) The organization of the human cerebral cortex estimated by intrinsic functional connectivity. Journal of neurophysiology 106:1125-1165. Yoshimura Y, Dantzker JL, Callaway EM (2005) Excitatory cortical neurons form fine-scale functional networks. Nature 433:868-873. 33
1902.06250
1
1902
2019-02-17T12:32:50
Afferent Fiber Activity-Induced Cytoplasmic Calcium Signaling in Parvalbumin-Positive Inhibitory Interneurons of the Spinal Cord Dorsal Horn
[ "q-bio.NC", "q-bio.TO" ]
Neuronal calcium (Ca2+) signaling represents a molecular trigger for diverse central nervous system adaptations and maladaptions. The altered function of dorsal spinal inhibitory interneurons is strongly implicated in the mechanisms underlying central sensitization in chronic pain. Surprisingly little is known, however, about the characteristics and consequences of Ca2+ signaling in these cells, including whether and how they are changed following a peripheral insult or injury and how such alterations might influence maladaptive pain plasticity. As a first step towards clarifying the precise role of Ca2+ signaling in dorsal spinal inhibitory neurons for central sensitization, we established methods for characterizing Ca2+ signals in genetically defined populations of these cells. In particular, we employed recombinant adeno-associated viral vectors to deliver subcellularly targeted, genetically encoded Ca2+ indicators into parvalbumin-positive spinal inhibitory neurons. Using wide-field microscopy, we observed both spontaneous and afferent fiber activity triggered Ca2+ signals in these cells. We propose that these methods may be adapted in future studies for the precise characterization and manipulation of Ca2+ signaling in diverse spinal inhibitory neuron subtypes, thereby enabling the clarification of its role in the mechanisms underlying pain chronicity and opening the door for possibly novel treatment directions.
q-bio.NC
q-bio
Title Afferent Fiber Activity-Induced Cytoplasmic Calcium Signaling in Parvalbumin-Positive Inhibitory Interneurons of the Spinal Cord Dorsal Horn Authors Anna M. Hagenston*, Sara Ben Ayed, Hilmar Bading Department of Neurobiology Interdisciplinary Center for Neurosciences (IZN) Heidelberg University INF 366 69120 Heidelberg Germany *Corresponding author Anna M. Hagenston: [email protected] Sara Ben Ayed: [email protected] Hilmar Bading: [email protected] 1 Abstract Neuronal calcium (Ca2+) signaling represents a molecular trigger for diverse central nervous system adaptations and maladaptions. The altered function of dorsal spinal inhibitory interneurons is strongly implicated in the mechanisms underlying central sensitization in chronic pain. Surprisingly little is known, however, about the characteristics and consequences of Ca2+ signaling in these cells, including whether and how they are changed following a peripheral insult or injury and how such alterations might influence maladaptive pain plasticity. As a first step towards clarifying the precise role of Ca2+ signaling in dorsal spinal inhibitory neurons for central sensitization, we established methods for characterizing Ca2+ signals in genetically defined populations of these cells. In particular, we employed recombinant adeno-associated viral vectors to deliver subcellularly targeted, genetically encoded Ca2+ indicators into parvalbumin-positive spinal inhibitory neurons. Using wide-field microscopy, we observed both spontaneous and afferent fiber activity triggered Ca2+ signals in these cells. We propose that these methods may be adapted in future studies for the precise characterization and manipulation of Ca2+ signaling in diverse spinal inhibitory neuron subtypes, thereby enabling the clarification of its role in the mechanisms underlying pain chronicity and opening the door for possibly novel treatment directions. Keywords inhibitory neuron, dorsal spinal cord, calcium, central sensitization, GCaMP3, parvalbumin Main text Changes in intracellular calcium (Ca2+) levels in neurons are intimately linked to mechanisms underlying adaptive plastic changes within the central nervous system. Ca2+ rises influence neuronal properties by diverse means, including modulation of ion channel activity or localization, regulation of neurotransmitter release, activation of Ca2+-dependent 2 enzymes, and induction or downregulation of gene expression. Dysregulation of intracellular Ca2+ levels and their stimulus-evoked changes are similarly implicated in several central nervous system pathologies; aberrant intraneuronal Ca2+ signaling is involved in processes as broad-ranging as excitotoxic neuronal death following stroke (1), impaired synaptic plasticity and memory formation in aging (2), and synapse loss in Alzheimer's disease (3), but also peripheral and central sensitization in chronic pain (4). Spinal central sensitization, defined as the aberrant amplification of noxious signal transmission in the spinal cord, is characterized both by the enhanced excitation of projection neurons and by a loss of inhibitory tone. Numerous mechanisms have been proposed to explain altered spinal inhibitory tone in central sensitization, including reduced excitation or excitability of inhibitory interneurons, degeneration of inhibitory neurons and/or their processes, reduced numbers of GABA- and glycinergic synaptic terminals, and diminished inhibitory neurotransmitter release (5, 6). Nonetheless, the precise molecular triggers underlying these alterations have not been elucidated. In light of its importance for regulating diverse neuronal functions, dysregulated intracellular Ca2+ signaling clearly represents a possible proximal trigger for altered inhibitory neuron function in central sensitization. Surprisingly little is known, however, about the characteristics and consequences of Ca2+ signaling in dorsal spinal inhibitory neurons, including whether it is altered in chronic pain states. We therefore aimed to establish methods for measuring Ca2+ signals in defined inhibitory neuron types and subcellular compartments. Here we present data demonstrating the use of these methods for characterizing cytoplasmic Ca2+ responses of parvalbumin (PV)-positive inhibitory neurons in the dorsal spinal cord to stimulation of sensory fiber afferents. We used the genetically encoded calcium indicator (GECI) GCaMP3, linked to a nuclear export signal (NES), to characterize cytoplasmic Ca2+ signals. While the maximum response amplitude and kinetics of GCaMP3 are inferior to those of, for example, GCaMP6f (7), this particular GECI has the advantage that its relatively high baseline fluorescence allows for 3 the easy identification of transduced cells. Thus, GCaMP3 is our preferred GECI for wide- field Ca2+ imaging in acute tissue preparations. We used recombinant adeno-associated viral (rAAV)-mediated gene delivery to express cytoplasmically-localized GCaMP3 (GCaMP3.NES) in PV-positive neurons of the spinal cord dorsal horn of PV-Cre mice. Specifically, we generated rAAV vectors containing a doubly floxed inverted open reading frame for GCaMP3.NES, which is expressed in a Cre recombinase-dependent manner under control of the human EF1α promoter (Fig. 1A). Initial tests carried out in primary hippocampal cultures co-infected (or not) with a second rAAV driving the expression of myc- tagged Cre recombinase -- detected immunocytochemically -- showed no GCaMP3.NES expression in the absence of Cre recombinase (data not shown). We injected rAAV vectors into the spinal parenchyma at spinal level L3/L4 and, after an incubation period of 3-4 weeks to allow for stable transgene expression, prepared acute transverse slices from the L3-L5 spinal cord with attached dorsal roots for Ca2+ imaging (Fig. 1B). Consistent with numerous studies demonstrating that PV-positive inhibitory neurons are localized in lamina III and make projections into lamina II of the rodent spinal cord dorsal horn (8, 9), we observed GCaMP3.NES fluorescence in multiple cell bodies and processes within laminae II-III of the dorsal spinal cord (Fig. 1C). Aβ, Aẟ, and C afferent fibers, which carry information related to innocuous touch, fast pain, and slow pain, respectively, were electrically stimulated with varying frequency and amplitude through a suction electrode attached to the dorsal root. The chosen stimulation frequencies (4, 20, and 100 Hz) were roughly tuned to the maximum following frequencies of each fiber class (10, 11), and a broad range of stimulation intensities (0.01-2.0 mA) was employed in order to progressively activate each fiber class (12). As we have previously reported for excitatory neurons in laminae I-II recorded under similar conditions (13), PV- positive inhibitory neurons produced graded Ca2+ responses to afferent fiber stimulation of increasing frequency and intensity (Fig. 1D,E; n=3-9 slices from 3-7 mice). PV-positive 4 inhibitory neurons in lamina III predominantly receive synaptic input from Aβ sensory afferents (8). In keeping with this feature, we observed no sharp rise in response amplitudes above Aẟ (~0.1 mA) or C (~0.5 mA) stimulation intensities for 20 Hz and 4 Hz stimulation trains, respectively. In most recorded slices (n=13/16 slices from 9 mice), PV-positive cell bodies and processes exhibited spontaneous Ca2+ rises in the absence of synaptic input from the periphery (Fig. 1D), suggestive of their involvement in a basally active and recurrent spinal network. We have not observed such spontaneous activity in dorsal spinal excitatory neurons (data not shown). The data we present here demonstrate that PV-positive inhibitory neurons in the dorsal spinal cord exhibit both afferent activity-evoked and spontaneous cytoplasmic Ca2+ rises. Following peripheral nerve injury, the number of synapses formed by these cells onto excitatory interneurons decreases (9). Consequently, innocuous tactile information impinging on laminae III and IV gains access to nociceptive circuits in laminae I/II, and mechanical allodynia arises (9). A possible molecular mechanism underlying inhibitory synapse loss is the impaired activity- and Ca2+-dependent expression of genes involved in inhibitory synapse maintenance (14). Thus, an examination of afferent activity-evoked Ca2+ transients following nerve injury -- embedded in a study employing gain- and loss-of-function tools specifically in PV-positive dorsal spinal neurons -- may provide important mechanistic insight into spinal disinhibition in neuropathic pain. We limited ourselves here to the description of Ca2+ signaling in one inhibitory neuron type and subcellular compartment. Given the expanding repertoire of available Cre mouse lines for defined populations of inhibitory neurons, the broad range of proven subcellular targeting signals for GECIs, and the growing number of genetically encoded calcium signaling inhibitors (e.g., CaMBP4, PV.NLS (13, 15)), we envision that our methods can be easily adapted for studies aimed at gaining a precise understanding of the characteristics and 5 consequences of spinal inhibitory neuronal Ca2+ in central sensitization and pain chronification. Extended Materials and Methods Mice Mice expressing the Cre recombinase in parvalbumin (PV)-positive inhibitory neurons (PV- Cre mice) were bred from founders originally obtained from Jackson Laboratories (JAX Stock #008069 (16)) and kindly provided by R. Kuner (Heidelberg University). All experimental procedures were approved by the local animal care and use committee (Regierungspräsidium Karlsruhe, Referat 35, Karlsruhe, Germany), project G-272/14, and carried out in accordance with ARRIVE guidelines and the EU Directive 2010/63/EU for animal experiments. Mice had access to food and water ad libitum and were housed on a 12 h light-dark cycle. Production and intra-spinal injection of rAAV vectors The coding sequence for GCaMP3.NES, targeted to the cytoplasm using a nuclear export signal (NES), was subcloned using standard procedures between the NheI and AscI restriction sites of the Cre recombinase-dependent AAV expression plasmid, pAAV-EF1a- double floxed-EYFP-WPRE-HGHpA, a gift from Karl Deisseroth (Addgene plasmid # 20296; http://n2t.net/addgene:20296; RRID:Addgene_20296). Recombinant serotype 1/2 adeno- associated viral vectors (rAAV) were produced and purified using previously described methods (17, 18) and injected into the lumbar spinal cord of adult PV-Cre mice. Briefly, mice were anesthetized with an intraperitoneal injection of sleep mix containing 0.5 mg/kg medetomidine, 5.0 mg/kg midazolam, and 0.5 mg/kg fentanyl. When the mice were no longer responsive to a foot pinch, an ~1 cm incision was made in the skin over the vertebral column, and the tissue between thoracic vertebrae T12 and T13 -- corresponding to spinal lumbar segments L3/L4 -- was dissected to reveal the dorsal spinal cord. The dura mater 6 was opened to improve needle penetration, and a 35 G beveled needle mounted in a 10 μl Hamilton syringe (World Precision Instruments) was slowly advanced into the spinal parenchyma. 500 nl of a 2:1 mixture of rAAV vector suspended in PBS and 20 % mannitol containing ~1011 virions was then injected at a rate of 100 nl/min using a microprocessor- controlled minipump (World Precision Instruments). The needle was left in place for 5 min after the injection was complete to allow the viral infusion to diffuse away from the injection site. The wound was subsequently covered with gelatin foam (Gelita Medical) and the skin closed with sutures. Animals received carprofen (5 mg/kg s.c.) prior to the first incision and buprenorphine (0.1 mg/kg s.c.) 1 h after subcutaneous administration of wake mix containing 0.75 mg/kg atipamezole, 0.5 mg/kg flumazenil and 1.2 mg/kg naloxone to minimize post- surgical pain. After the procedure, mice were returned their home cage, which was placed on a heating plate for 24 h, and provided with wet food pellets. As we have demonstrated in the past (13), this viral delivery method generally results in transgene expression spreading from L2 to the border of S1 of the mouse spinal cord with a maximum in L3-L5, the area of the spinal cord that innervates the hindpaw. Ca2+ imaging and dorsal root stimulation Three to four weeks following rAAV injection surgery, acute spinal cord slices with attached dorsal roots from spinal cord segments L3-L5 of rAAV-injected mice were prepared as described previously (19). Briefly, mice were anesthetized with sodium pentobarbital (320 mg/kg, i.p.) and perfused with freshly prepared, ice-cold NMDG slicing solution containing, in mM, 93 HCl, 2.5 KCl, 1.2 NaH2PO4, 3 sodium pyruvate, 20 HEPES, 30 NaHCO3, 25 D- glucose, 5 L-ascorbic acid, 10 N-acetyl-L-cysteine, 2 thiourea, 10 MgSO4, 0.5 CaCl2, adjusted to a final osmolarity of 300-305 mOsm and pH of 7.35, and bubbled with 95 % O2/5 % CO2. Afterwards, the spinal cord was isolated, the dura mater carefully removed, and the ventral roots resected. The spinal cord was then mounted into a groove within a block of 1.5 7 % agarose in PBS and secured with cyanoacrylate adhesive to the stage of a Microm HM650V (Thermo Scientific) vibratome containing NMDG slicing solution bubbled with 95 % O2/5 % CO2 and cooled to 1-4 °C. Slices (~330 µm) were prepared using double-edge coated blades (Science Services, washed version, #7200-WA) and then placed into NMDG slicing solution bubbled with 95 % O2/5 % CO2 and warmed to 34 °C for 10 min. Slices were subsequently transferred to extracellular recording solution containing, in mM, 125 NaCl, 10 D-glucose, 25 NaHCO3, 2.5 KCl, 1.25 NaH2PO4, 1.2 MgSO4, 2 CaCl2, adjusted to a final osmolarity of 300-305 mOsm and pH 7.35, bubbled with 95 % O2/5 % CO2 and warmed to 34 °C, and then allowed to cool to room temperature for a recovery period of ≥ 1 h. For Ca2+ imaging, slices were transferred to a recording chamber (PM-1, Warner Instruments, Hamden, CT, USA or PC-R, Siskiyou, OR, USA) perfused (~3 ml/min) with 95 % O2/5 % CO2-bubbled extracellular recording solution warmed with an in-line heater to 34 °C. Slices were secured with a platinum ring with nylon strings. Dorsal roots were sucked into a glass stimulation electrode, the tip of which had been dipped into Vaseline to improve the electrical seal around the root, and stimulated with 1 s trains of electrical pulses (0.1 ms pulse width, 100 Hz and 20 Hz stimuli; 1 ms pulse width, 4 Hz stimuli) using a stimulus isolator (World Precision Instruments, A365) controlled by Clampex 10.3 Software (Molecular Devices). Ca2+ responses from GCaMP3.NES-labeled cells and their processes in laminae II-III were imaged as described in detail previously (20). GCaMP3.NES was excited using a CoolLED light source (480±10 nm), and fluorescence emission (530±20 nm) acquired at a rate of 2 Hz with a cooled camera (Photometrics Coolsnap HQ, Roper Scientific) with 2x2 binning through a 20x water-immersion objective (XLM PlanFluor 0.95W, Olympus) on an upright microscope (Olympus BX51WI) connected to a software interface (Metafluor, Universal Imaging Systems and Molecular Devices). Overview images were taken with a 4x air objective (PlanCN, Olympus). 8 Data analysis Image sequences were imported into ImageJ (Fiji RRID:SCR_002285) and, when necessary, processed with the "Template Matching" plugin to correct for slice movement during the recording. ROIs were manually drawn around single cells and larger regions encompassing nearly the entire field of view within laminae II/III containing fluorescently labeled cells and processes, and fluorescence intensity changes measured over time. Further data analysis was carried out using Igor Pro (WaveMetrics, RRID:SCR_000325). Fluorescence intensity changes were expressed as the percent change with respect to baseline in the 15 s prior to the onset of the stimulation train (% ΔF/F) and quantified using the peak amplitude above baseline. Data were plotted as mean ± SEM using GraphPad Prism 6 (RRID:SCR_002798). Post-imaging analysis of GCaMP3.NES expression Following Ca2+ imaging, acute spinal cord slices were fixed in 4 % paraformaldehyde in PBS for 2 h than them placed in 30 % sucrose in PBS containing 0.04 % Thimerosal (Sigma- Aldrich) for a minimum of 24 h. Tissue slices were embedded in tissue freezing medium (Leica), and 20 μm sections cut using a cryotome (CM 1950, Ag Protect, Leica) and mounted directly onto SuperFrost Plus slides (Thermo Scientific). Slides were coverslipped with Mowiol 4-88 (Calbiochem) medium containing 2 µg/ml Hoechst 33258 (Serva, Cat # 15090) as a nuclear counterstain. Images of the dorsal horn were obtained with an upright wide- field microscope (DM IRBE, Leica) and 20x objective and processed using Adobe Photoshop CS4. List of abbreviations Ca2+: calcium GECI: genetically encoded calcium indicator 9 NES: nuclear export signal PV: parvalbumin rAAV: recombinant adeno-associated virus Declarations Ethics approval This study was approved by the local animal care and use committee (Regierungspräsidium Karlsruhe, Referat 35, Karlsruhe, Germany), project G-272/14. All animal experiments were designed and carried out in accordance with ARRIVE guidelines and the EU Directive 2010/63/EU for animal experiments. Availability of data and materials The data sets analyzed during the current study are available from the corresponding author upon reasonable request. Consent for publication Not applicable Competing interests The authors declare that they have no competing interests. Funding This study was funded by the German Research Foundation (SFB1158, grant A05). Acknowledgements 10 We would like to thank Rohini Kuner (Heidelberg University) for providing PV-Cre mice for use in this study. Authors' contributions AMH and HB conceived the project, designed the study, and wrote the paper. AMH and SK performed the experiments. AMH and HB analyzed and interpreted the data. All authors read and approved the final manuscript. References 1. Hardingham GE, Bading H. Synaptic versus extrasynaptic NMDA receptor signalling: implications for neurodegenerative disorders. Nat Rev Neurosci. 2010;11:682-96. 2. Kumar A, Bodhinathan K, Foster TC. Susceptibility to Calcium Dysregulation during Brain Aging. Front Aging Neurosci. 2009;1:2. 3. Bezprozvanny I, Mattson MP. Neuronal calcium mishandling and the pathogenesis of Alzheimer's disease. Trends Neurosci. 2008;31(9):454-63. 4. Hagenston AM, Simonetti M. Neuronal calcium signaling in chronic pain. Cell Tissue Res. 2014;357(2):407-26. 5. Alles SRA, Smith PA. Etiology and Pharmacology of Neuropathic Pain. Pharmacol Rev. 2018;70(2):315-47. 6. Zeilhofer HU, Wildner H, Yevenes GE. Fast synaptic inhibition in spinal sensory processing and pain control. Physiol Rev. 2012;92(1):193-235. 7. Chen TW, Wardill TJ, Sun Y, Pulver SR, Renninger SL, Baohan A, et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature. 2013;499(7458):295-300. 8. Todd AJ. Identifying functional populations among the interneurons in laminae I-III of the spinal dorsal horn. Mol Pain. 2017;13:1744806917693003. 11 9. Petitjean H, Pawlowski SA, Fraine SL, Sharif B, Hamad D, Fatima T, et al. Dorsal Horn Parvalbumin Neurons Are Gate-Keepers of Touch-Evoked Pain after Nerve Injury. Cell Rep. 2015;13(6):1246-57. 10. Daniele CA, MacDermott AB. Low-threshold primary afferent drive onto GABAergic interneurons in the superficial dorsal horn of the mouse. J Neurosci. 2009;29(3):686-95. 11. Waddell PJ, Lawson SN. Electrophysiological properties of subpopulations of rat dorsal root ganglion neurons in vitro. Neuroscience. 1990;36(3):811-22. 12. Sdrulla AD, Xu Q, He SQ, Tiwari V, Yang F, Zhang C, et al. Electrical stimulation of low-threshold afferent fibers induces a prolonged synaptic depression in lamina II dorsal horn neurons to high-threshold afferent inputs in mice. Pain. 2015;156(6):1008-17. 13. Simonetti M, Hagenston AM, Vardeh D, Freitag HE, Mauceri D, Lu J, et al. Nuclear calcium signaling in spinal neurons drives a genomic program required for persistent inflammatory pain. Neuron. 2013;77(1):43-57. 14. Sun X, Lin Y. Npas4: Linking Neuronal Activity to Memory. Trends Neurosci. 2016;39(4):264-75. 15. Mauceri D, Hagenston AM, Schramm K, Weiss U, Bading H. Nuclear Calcium Buffering Capacity Shapes Neuronal Architecture. J Biol Chem. 2015;290(38):23039-49. 16. Hippenmeyer S, Vrieseling E, Sigrist M, Portmann T, Laengle C, Ladle DR, et al. A developmental switch in the response of DRG neurons to ETS transcription factor signaling. PLoS Biol. 2005;3(5):e159. 17. McClure C, Cole KL, Wulff P, Klugmann M, Murray AJ. Production and titering of recombinant adeno-associated viral vectors. J Vis Exp. 2011(57):e3348. 18. Zhang SJ, Steijaert MN, Lau D, Schutz G, Delucinge-Vivier C, Descombes P, et al. Decoding NMDA receptor signaling: identification of genomic programs specifying neuronal survival and death. Neuron. 2007;53(4):549-62. 12 19. Luo C, Seeburg PH, Sprengel R, Kuner R. Activity-dependent potentiation of calcium signals in spinal sensory networks in inflammatory pain states. Pain. 2008;140(2):358-67. 20. Bengtson CP, Freitag HE, Weislogel JM, Bading H. Nuclear calcium sensors reveal that repetition of trains of synaptic stimuli boosts nuclear calcium signaling in CA1 pyramidal neurons. Biophys J. 2010;99(12):4066-77. 13 Figures Figure 1. Afferent activity-triggered and spontaneous Ca2+ signaling in dorsal spinal inhibitory interneurons. (A) We generated serotype 1/2 rAAV vectors containing a doubly floxed inverted open reading frame coding for the cytoplasmically-localized GECI, GCaMP3.NES, expressed in a Cre recombinase-dependent manner under control of the EF1α promoter. (B) rAAV vectors were injected into the spinal parenchyma of adult PV-Cre mice. Acute slices with attached dorsal roots from the L3-L5 spinal cord were prepared 3-4 weeks later. Sensory afferent activity was evoked by suction electrode stimulation of the attached dorsal root. Recordings of Ca2+ responses from PV-positive neurons were made in laminae II-III of the spinal cord dorsal horn. (C) Analysis of GCaMP3.NES expression in the spinal cord dorsal horn. GCaMP3.NES was observed in the extranuclear soma and processes of cells residing in lamina III and extending processes into lamina II. The upper border of lamina I is marked with a dashed line. Scale bar, 100 μm. (D) Top, representative image from a Ca2+ imaging experiment in which multiple GCaMP3.NES-positive neurons 14 could be visualized. Bottom, changes in cytoplasmic Ca2+ levels were visualized using the normalized fluorescence intensity in regions of interest laid over individual neurons (red, green, and blue traces; corresponding the colored arrows in the image above) and the bulk tissue (black traces; corresponding to the large white ellipse in the image above) during 100 Hz dorsal root stimulation (0.2-2.0 mA). Both stimulus-evoked responses and spontaneous activity with highly variable amplitude and kinetics were observed. (E) Quantification of Ca2+ rises in the bulk tissue triggered by 1 s, 0.01-2.0 mA stimulation trains delivered at 4, 20, and 100 Hz (n = 3-9 slices from 3-7 animals). The amplitudes of evoked Ca2+ transients varied positively with both stimulus frequency and intensity. 15
1602.07389
3
1602
2017-07-07T20:22:20
Capturing the dynamical repertoire of single neurons with generalized linear models
[ "q-bio.NC" ]
A key problem in computational neuroscience is to find simple, tractable models that are nevertheless flexible enough to capture the response properties of real neurons. Here we examine the capabilities of recurrent point process models known as Poisson generalized linear models (GLMs). These models are defined by a set of linear filters, a point nonlinearity, and conditionally Poisson spiking. They have desirable statistical properties for fitting and have been widely used to analyze spike trains from electrophysiological recordings. However, the dynamical repertoire of GLMs has not been systematically compared to that of real neurons. Here we show that GLMs can reproduce a comprehensive suite of canonical neural response behaviors, including tonic and phasic spiking, bursting, spike rate adaptation, type I and type II excitation, and two forms of bistability. GLMs can also capture stimulus-dependent changes in spike timing precision and reliability that mimic those observed in real neurons, and can exhibit varying degrees of stochasticity, from virtually deterministic responses to greater-than-Poisson variability. These results show that Poisson GLMs can exhibit a wide range of dynamic spiking behaviors found in real neurons, making them well suited for qualitative dynamical as well as quantitative statistical studies of single-neuron and population response properties.
q-bio.NC
q-bio
Capturing the Dynamical Repertoire of Single Neurons with Generalized Linear Models Alison I. Weber1 & Jonathan W. Pillow2,3 1Graduate Program in Neuroscience, University of Washington, Seattle, WA, USA. 2Princeton Neuroscience Institute, Princeton University, Princeton, NJ, USA. 3Dept. of Psychology, Princeton University, Princeton, NJ, USA. Keywords: point process; generalized linear model (GLM); Izhikevich model; spike timing; variability Abstract A key problem in computational neuroscience is to find simple, tractable models that are never- theless flexible enough to capture the response properties of real neurons. Here we examine the capabilities of recurrent point process models known as Poisson generalized linear models (GLMs). These models are defined by a set of linear filters, a point nonlinearity, and condi- tionally Poisson spiking. They have desirable statistical properties for fitting and have been widely used to analyze spike trains from electrophysiological recordings. However, the dynam- ical repertoire of GLMs has not been systematically compared to that of real neurons. Here we show that GLMs can reproduce a comprehensive suite of canonical neural response be- haviors, including tonic and phasic spiking, bursting, spike rate adaptation, type I and type II excitation, and two forms of bistability. GLMs can also capture stimulus-dependent changes in spike timing precision and reliability that mimic those observed in real neurons, and can exhibit varying degrees of stochasticity, from virtually deterministic responses to greater-than-Poisson variability. These results show that Poisson GLMs can exhibit a wide range of dynamic spiking behaviors found in real neurons, making them well suited for qualitative dynamical as well as quantitative statistical studies of single-neuron and population response properties. 7 1 0 2 l u J 7 ] . C N o i b - q [ 3 v 9 8 3 7 0 . 2 0 6 1 : v i X r a 1 Introduction Understanding the dynamical and computational properties of neurons is a fundamental chal- lenge in cellular and systems neuroscience. A wide variety of single-neuron models have been proposed to account for neural response properties. These models can be arranged along a complexity axis ranging from detailed, interpretable, biophysically accurate models to simple, tractable, reduced functional models. Detailed Hodgkin-Huxley style models, which sit at one end of this continuum, provide a biophysically detailed account of the conductances, currents, and channel kinetics governing neural response properties [21]. These models can account for the vast dynamical repertoire of real neurons, but they are often unwieldy for theoretical analyses of neural coding and computation. This motivates the need for simplified models of neural spike responses that are tractable enough for mathematical, computational, and statis- tical analyses. A variety of simplified dynamical models have been proposed to serve the need for math- ematically tractable models, including the integrate-and-fire model, Fitzhugh-Nagumo, Morris- Lecar, and Izhikevich models [9, 33, 32, 22, 4]. Generally, these models aim to reduce the biophysically detailed descriptions of realistic neurons to systems of differential equations with fewer variables and/or simplified dynamics. The one-dimensional integrate-and-fire model is arguably the simplest of these, and the simplest to analyze mathematically, but it fails to cap- ture many of the response properties of real neurons. The two-dimensional Izhikevich model, by contrast, was specifically formulated to retain the rich dynamical repertoire of more com- plex, biophysically realistic models [23]. An alternative to a mathematical notion of simplicity is the statistical property of being tractable for fitting from intracellular or extracellular physiological recordings. One well-known statistical model that satisfies this desideratum is the recurrent linear-nonlinear Poisson model, commonly referred to in the neuroscience literature as the generalized linear model (GLM) [48, 40]. GLMs are closely related to generalized integrate-and-fire models such as the spike- response model, which has linear dynamics but incorporates spike-dependent feedback to capture the nonlinear effects of spiking on neural membrane potential and subsequent spike generation [13, 26, 24, 39, 16]. In fact, a variant of the spike response model that incorporates noise into the spike threshold is mathematically equivalent to the models we study here [15, 12, 25, 14]. GLMs are popular for characterizing neural responses in reverse-correlation or white-noise experiments, due to the tractability of likelihood-based fitting methods. Recent work has shown that GLMs can capture the detailed statistics of spiking in single and multi- neuron recordings from a variety of brain areas [40, 2, 5, 50, 31, 41]. While several studies have shown that GLMs can successfully recapitulate various re- sponse properties of biological or simulated neurons, here, we provide a more systematic study of the dynamical repertoire of the GLM . We study this issue by fitting GLMs to data from simulated neurons exhibiting a number of complex response properties. We show that GLMs can reproduce a remarkably rich set of dynamical behaviors, including tonic and phasic spiking, bursting, spike rate adaptation, type I and type II excitation, and two different forms of bistability. Furthermore, GLMs can exhibit stimulus-dependent degrees of spike timing preci- sion and reliability [29], and mimic a recently reported form of greater-than-Poisson variability [17]. 2 2 Models of dynamical behaviors 2.1 Izhikevich model First, we will examine whether generalized linear models can reproduce a suite of canonical spiking behaviors exhibited by the well-known Izhikevich model [22, 23]. The Izhikevich model is a biophysically-inspired model of intracellular membrane potential defined by a two-variable system of ordinary differential equations governing membrane potential v(t) and a recovery variable u(t): v = 0.04v2 + 5v + 140 − u + I(t) u = a(bv − u) with spiking and voltage-reset governed by the boundary condition: if v(t) ≥ 30, "spike" and set (cid:26) v(t+) = c u(t+) = u(t) + d, (1) (2) (3) where I(t) is injected current, t+ denotes the next time step after t, and parameters (a, b, c, d) determine the model's dynamics. Different settings of these parameters lead to qualitatively different spiking behaviors, as shown in [23]. We focus on this model because of its demon- strated ability to produce a wide range of response properties exhibited by real neurons. (See Table 1 for parameter values used in this study and Methods for simulation details.) 2.2 Generalized linear model (GLM) The GLM is a regression model typically used to characterize the relationship between exter- nal or internal covariates and a set of recorded spike trains. In systems neuroscience, the label "GLM" often refers to an autoregressive point process model, a model in which linear functions of stimulus and spike history are nonlinearly transformed to produce the spike rate or conditional intensity of a Poisson process [48, 40]. The GLM is parametrized by a stimulus filter (cid:126)k, which describes how the neuron integrates an external stimulus, a post-spike filter (cid:126)h, which captures the influence of spike history on the current probability of spiking, and a scalar µ that determines the baseline spike rate. (See Figure 1.) The outputs of these filters are summed and passed through a nonlinear function f to determine the conditional intensity λ(t): λ(t) = f ((cid:126)k · (cid:126)x(t) + (cid:126)h · (cid:126)yhist (t) + µ), (4) where (cid:126)x(t) is the (vectorized) spatio-temporal stimulus, (cid:126)yhist(t) is a vector representing spike history at time t, and f is a nonlinear function that ensures the spike rate is non-negative. Spikes are generated according to a conditionally Poisson process [37, 6], so spike count y(t) in a time bin of size ∆ is distributed according to a Poisson distribution: P (y(t)λ(t)) = 1 y(t)! ∆λ(t) e−∆λ(t). (5) (cid:17)y(t) (cid:16) 3 Figure 1: Schematic of the generalized linear model. A: The stimulus filter (cid:126)k operates linearly on the stimulus x(t), is combined with input from the post-spike filter (cid:126)h and mean input level µ. This combined linear signal passes through a point nonlinearity f (·), whose output drives spiking via a conditionally Poisson process. B: An equivalent view of the GLM, which emphasizes the dependencies between a particular time window of stimulus and spike history and conditional intensity λ(t), which governs the probability of a spike in the current time bin (dark gray box). In this study we set f to be exponential, although similar properties can be obtained with other nonlinearities such as the soft-rectification function. Unlike classical deterministic models like Hodgkin-Huxley and integrate-and-fire, the GLM is fundamentally stochastic due to the assumption of conditionally Poisson spiking. However, this stochasticity is helpful for fitting purposes because it assigns graded probabilities to firing events and allows for likelihood-based methods for parameter fitting [35, 39]. In fact, the Poisson GLM comes with a well-known guarantee that the log-likelihood function is concave for suitable choices of nonlinearity f [34]. This means we can be assured of approaching a global optimum of the likelihood function via gradient ascent, for any set of stimuli and spike trains (barring any numerical issues that may complicate achieving the actual maximum for certain datasets, cf. [53]). This guarantee does not hold for stochastic formulations of most nonlinear biophysical models, including the Izhikevich model. Moreover, despite its stochasticity, the GLM can produce highly precise and repeatable spike trains in certain parameter settings, as we will demonstrate below. 3 GLMs capture a wide array of complex dynamical behav- iors We fit GLMs to data simulated from Izhikevich neurons set up to exhibit a range of different qualitative response behaviors. In the following, we describe these behaviors in detail, be- ginning with simpler behaviors, such as tonic spiking and bursting (which have already been demonstrated in previous work, e.g., [15, 25]) in order to build intuition for the GLM's basic capabilities, and then move on to more complex behaviors (such as bistability) and questions of spike timing reliability and precision. 4 generalized linear model (GLM)equivalent diagramABstimulusspike traintimestimulus filternonlinearitypost-spike filterstochasticspiking+ 3.1 Tonic spiking We first examined an Izhikevich neuron tuned to exhibit tonic spiking (Figure 2A-B; see Table 1 for parameters). When presented with a step input current, the Izhikevich neuron responds with a few high-frequency spikes and then settles into a regular firing pattern that persists for the duration of the step (Figure 2B). This response pattern resembles that of a deterministic Hodgkin-Huxley or integrate-and-fire neuron, albeit with an added transient burst of spikes at stimulus onset. We simulated the Izhikevich neuron's response to a series of step currents and used the resulting training data to perform maximum-likelihood fitting of the GLM parameters {(cid:126)k,(cid:126)h, µ}. The estimated stimulus filter (cid:126)k is biphasic (Figure 2C), resulting in a large transient response to stimulus onset, and has a positive integral, ensuring a sustained positive response to a Figure 2: Tonic spiking behavior. A: A step current stimulus. B: Voltage response of the simulated Izhikevich neuron. C: The fitted GLM stimulus filter (cid:126)k has a biphasic shape that gives the model a vigorous response to stimulus onset and a net positive response to a sus- tained input. D: The fitted GLM post-spike filter (cid:126)h has a negative lobe that imposes strong refractoriness on a timescale of ≈50 ms. E: Stimulus (blue) and post-spike (red) filter outputs during simulated response of the fitted GLM to the stimulus shown above on a single trial. F: The summed filter outputs are passed through an exponential nonlinearity to determine the conditional intensity λ(t), shown here for a single trial. G: Spike train of the Izhikevich neuron (black) and simulated repeats of the GLM (gray). GLM spike responses are slightly different on each trial, due to the stochasticity of spike generation, but reproduce Izhikevich model spike times with high precision. 5 Izhikevich neuron responseA!EFGinputmVfilter outputsresponsesGLM repeatsconditional!intensitytime (ms)stimulusC!time (ms)(sp/s)Izhikevich = -35fitted GLM parametersBDstimulus filterpost-spike filter outputoutput02004006001201010106-200002000-80-30200510-100-500-0.500.511.5050100-1000-750-500-2500 current step (Figure 2E, blue trace). The estimated post-spike filter (cid:126)h (Figure 2D), by contrast, has a large negative lobe that provides recurrent inhibition after every spike, enforcing a strong relative refractory period. The stimulus filter and post-spike filter output (shown together for a single trial in Figure 2E) are summed together and exponentiated to obtain the conditional intensity λ(t) (Figure 2F), also known as the instantaneous spike rate. For this stimulus, the intensity rises very quickly once (cid:126)h decays, which occurs approxi- mately 50 ms after the previous spike. Note that the output of the stimulus filter is identical on each trial, whereas the the output of the post-spike filter varies from trial to trial because of variability in the exact timing of spikes. However, because the rising phase of the condi- tional intensity is so rapid, spiking is virtually certain within a small time window sitting at a fixed latency after the previous spike time. The combination of strong excitatory drive from the stimulus filter and strong suppressive drive from the post-spike filter produces precisely timed spikes across trials, allowing the GLM to closely match the deterministic firing pattern of the Izhikevich neuron (Figure 2G). 3.2 Bursting We next examined multi-spike bursting, a more complex temporal response pattern that re- quires dependencies beyond the most recent interspike interval. Once again, we simulated responses from an Izhikevich neuron tuned to exhibit tonic bursting (Figure 3A-B) and used the resulting data to fit GLM parameters (Figure 3C-D). The estimated stimulus filter (cid:126)k is bipha- sic with a larger positive than negative lobe, which drives rapid spiking at stimulus onset and generates sustained drive during an elevated stimulus (Figure 3E, blue trace). The post-spike filter (cid:126)h has an immediate negative component that creates a relative refractory period after each spike, and an even more negative mode after a latency of ≈40 ms; the accumulation of these negative components over multiple spikes gives rise to a sustained suppression of activity between bursts. The GLM captures the bursting behavior of the Izhikevich neuron with high precision, in- cluding the fact that the first burst after stimulus onset contains a different pattern of spikes than subsequent bursts. This difference arises from precise interactions between the stimulus and post-spike filter outputs. During the first burst, fast spiking arises from an interplay be- tween monotonically increasing stimulus filter output (Figure 3E, blue) and tonic decrements induced by the post-spike filter after each spike (Figure 3E, red). After each spike, the post- spike filter reduces the conditional intensity by a fixed decrement, but this decrement is soon overwhelmed by the rising wave of input from the stimulus filter, which creates a rapid rise and fall of the conditional intensity time-locked to each Izhikevich neuron spike time. The pattern continues until accumulated contributions from the delayed negative lobes of post-spike filter overwhelm those from the stimulus filter and the burst terminates. Subsequent bursts are gov- erned by a somewhat different interplay between stimulus and post-spike filter outputs: bursts sit on a rising phase of the conditional intensity due to the removal of suppression from the previous burst. This rise is more gradual than the drive induced by stimulus onset, and results in bursts with longer inter-spike intervals and fewer spikes per burst, but the resulting spike pattern is nonetheless captured with high precision and reliability from trial to trial. 6 Figure 3: Bursting behavior. A: Step current stimulus. B: Voltage response of Izhikevich neu- ron. C: Fitted GLM stimulus filter. D: Fitted GLM post-spike filter, which creates refractoriness on short timescales (within each burst) due to instantaneous depolarization following a spike. The large negative lobe ≈25-50 ms after a spike terminates bursting and strongly suppresses firing between bursts. E: Stimulus (blue) and post-spike (red) filter outputs for simulated re- sponse of the GLM to step current shown above on a single trial. F: Output of the nonlinearity (conditional intensity) λ(t) on a single trial. G: Spike train of Izhikevich neuron (black) and simulated repeats of the fitted GLM (gray). 3.3 Bistability Bistability refers to the phenomenon in which there are multiple stable response modes for a single input condition. A common form of bistability observed in real neurons is the ability to inhabit either a tonically active state or a silent state for a given level of current injection. The Izhikevich model can exhibit this form of bistability, wherein a brief positive current pulse is sufficient to kick it between states: the neuron can inhabit a silent state in the absence of stimulation, but a brief positive current pulse kicks it into a tonically active state, and an appropriately-timed positive pulse kicks it back to the silent state (Figure 4A-B). We fit a GLM to spike trains simulated from such a bistable Izhikevich neuron and found that the fitted GLM can capture bistable behavior of the original model with high accuracy (Figure 4C-G). When stimulated during the silent state, the GLM emits a spike due to the positive output of the stimulus filter (Figure 4E, blue), and tonic firing ensues due to a positive lobe in the post-spike filter that causes self-excitation at a fixed latency of approximately 10 ms after the previous spike (Figure 4E). The GLM returns from the active state to the silent state when a positive stimulus pulse synchronizes negative lobes of the stimulus and post-spike 7 3Izhikevich neuron responseA!EFGinputmVfilter outputsresponsesGLM repeatsconditional!intensitytime (ms)stimulusC!time (ms)(sp/s)Izhikevich = -121fitted GLM parametersBDstimulus filterpost-spike filter outputoutput050-200-150-100-500-100-500-0.100.10.20.30.4010-80-3020-1000-5000500106010100100200300400100 filters. Only the combination of these two negative drives is strong enough to shut off spiking; without the negative drive created by previous spikes (appearing in the post-spike filter at a latency of 15 ms after a spike), suppression from the negative lobe of the stimulus filter is not strong enough to prevent spiking. As with the tonic spiking neuron discussed above, the interaction between the stimulus and post-spike filters generates rapid rises in the conditional intensity (Figure 4F), leading to precisely timed spikes that mimic those of the deterministic Izhikevich neuron (Figure 4G). This Izhikevich neuron (with the same parameters) can also exhibit a second form of bista- bility, in which the return to the silent state from the active state is induced by a negative instead of a positive current pulse. We performed a similar fitting exercise and found that the GLM is also able to reproduce this behavior. Firing is initiated and maintained by a similar mechanism as the first form of bistability, but firing offset occurs due to the fact that a negative stimulus pulse creates immediate negative output from the stimulus filter, which suppresses firing during the time when a spike would have occurred due to spike-history filter input. Tonic firing is extinguished more rapidly in this second form of bistability than the first (see Figure 6 Figure 4: Bistable responses. A: Stimulus consisting of two brief positive current pulses. B: Voltage response of Izhikevich neuron, which exhibits bistability. The first pulse initiates a tonic spiking mode and the second pulse (precisely timed to the phase of the spike response) terminates it, returning to a quiescent mode. C: Fitted GLM stimulus filter, which provides a biphasic impulse response. D: Fitted post-spike filter, which imposes a refractory period of ≈4 ms and gives increased probability of firing ≈5-10 ms after each spike. E: GLM stimulus (blue) and post-spike (red) filter outputs on a single trial. F: Output of the nonlinearity gives the conditional intensity for a single trial. G: Spike train of Izhikevich neuron (black) and simulated repeats of the fitted GLM (gray). 8 Izhikevich neuron responseA!EFGinputmVfilter outputsresponsesGLM repeatsconditional!intensitytime (ms)stimulusC!time (ms)(sp/s)Izhikevich = -204fitted GLM parametersBDstimulus filterpost-spike filter outputoutput0601010103-60-50-40-80-3020-400-200050100-100-500-0.0200.0201020-400-2000 below). 3.4 Type I and type II firing Neurons have been classified as exhibiting either type I or type II dynamics based on the shape of their firing rate vs. intensity (F-I) curve. Type I neurons can fire at arbitrarily low rates for low levels of injected current, whereas type II neurons have discontinuous F-I curves that arise from an abrupt transition from silence to a finite non-zero firing rate as the level of injected current increases [20]. We simulated Izhikevich neurons that exhibit each of these response types using published parameters. Inputs consisted of 500 ms current steps of varying amplitude. The resulting F-I curves for the Izhikevich type I and type II neurons are shown in black in Figure 5A and Figure 5B, respectively. We fit GLMs using data from each Izhikevich neuron and found that the fitted GLMs capture the two response types with high temporal precision. The corresponding F-I curves are shown in gray in Figure 5 and accurately mimic the behaviors of the Izhikevich neuron. Similar F-I curves have been demonstrated previously in [15] and [31]. The only discrepancy between the Izhikevich and GLM neurons occurs for the type II cell at input amplitudes near the Izhikevich neuron's threshold. On some trials when the input amplitude falls below this threshold, the GLM jumps into a a tonic firing state for the duration Figure 5: Type I and type II firing curves. A: Top: Example responses of a GLM exhibiting type I firing behavior. The spike rate increases continuously from zero in response to current steps of increasing amplitude. Bottom: F-I curve for a type I Izhikevich neuron (black) and corresponding GLM (gray). For the GLM, responses are plotted for five repetitions of each input amplitude. B: Similar plots for type II firing behavior, characterized by a discontinuous jump from zero to a finite spike rate in responses to current steps of increasing amplitude. 9 15202530010203040input amplitudefiring rate (Hz) !! IzhikevichGLM0.20.40.6010203040input amplitudefiring rate (Hz)ABType IType II of the stimulus. Similarly, on some trials when the input amplitude falls above this threshold, the GLM fails to initiate firing. This is unsurprising given the stochastic nature of the GLM. Importantly, the GLM never fires at a low rate, but rather abruptly transitions from no firing to firing at a baseline level of ≈25 Hz, reflecting type II behavior. 3.5 Additional behaviors We fit GLMs to every dynamical behavior considered in [23] with the exception of purely sub- threshold behaviors, since GLM fitting uses spike trains and does not consider sub-threshold responses. The full suite of behaviors is shown in Figure 6, with responses of the Izhikevich neurons in black and spike responses of the GLM in gray. This list includes tonic and phasic spiking, tonic and phasic bursting, mixed mode firing, spike frequency adaptation, type I and type II excitability, two different forms of bistability, and several others that depend primarily on the shape of stimulus filter. Several additional behaviors that can be captured by a GLM are not depicted in Figure 6 as they can be achieved by a trivial manipulation of the stimulus filter; for example, inhibition-induced bursting can be achieved by simply flipping the sign of the stimulus filter for the bursting neuron shown in Figure 6C. Previous work has shown the Izhikevich neuron to be capable of producing 18 distinct spiking behaviors [23], and we found that all can also be produced by a GLM. 3.6 Systematic variation of filter amplitudes We next considered what happens to the behaviors produced by a GLM as some aspect of the filters is systematically varied. To do so, we created stimulus and post-spike filters composed by linear combinations of two basis filters, and then systematically varied the amplitude of one basis filter while holding the other fixed. (See Methods for details.) Figure 7 shows the phase space of qualitative spiking behaviors obtained at different points in this 2D filter space. When the stimulus filter has a strongly negative component (center panel, bottom), a pos- itive stimulus pulse does not produce enough driving force to cause the neuron to spike at all (quiescent). As the amplitude of this component of the stimulus filter is increased, the neuron receives stronger and stronger input and is driven first to spike once or twice (phasic spiking), and eventually to emit a burst of spikes (phasic bursting). The stimulus filter largely drives changes between these behaviors, with the additional detail that a strongly negative post- spike filter component is able to inhibit a burst that would otherwise occur (upper left corner of "phasic spiking" region). As the stimulus filter component becomes still more positive and the stimulus filter transitions from being biphasic to more monophasic, it produces a positive driving force for the duration of the stimulus step, rather than just the onset. This causes the neuron to fire for the duration of the step. Importantly, the post-spike filter here determines the nature of this sustained firing. A post-spike filter that is purely negative beginning at short timescales (top, left side) mimics a relative refractory period, inhibiting additional spikes for a short window following each elicited spike and resulting in tonic spiking. If, on the other hand, the post-spike filter is only weakly negative at short timescales while being more strongly negative at longer timescales, this creates multiple timescales in the neuron's response (top, right side). At short timescales, 10 Figure 6: Suite of dynamical behaviors of Izhikevich and GLM neurons. Each panel, top to bottom: stimulus (blue), Izhikevich neuron response (black), GLM responses on five trials (gray), stimulus filter (left, blue), and post-spike filter (right, red). Black line in each plot indicates a 50 ms scale bar for the stimulus and spike response. (Differing timescales reflect timescales used for each behavior in original Izhikevich paper [23]). Stimulus filter and post- spike filter plots all have 100 ms duration. there is little inhibition from each spike (beyond the absolute refractory period), so additional spikes may occur. Over longer timescales, inhibition is accumulated over multiple spikes, which eventually shuts off spiking. After a brief window of no spikes, the inhibition is relaxed and spiking commences again until enough inhibition is accumulated to shut spiking off. This cycle results in tonic bursting for the duration of the step. In the extreme case where the post-spike filter is actually biphasic (top, far right), each spike promotes additional spikes on short timescales, leading to highly regular timing of spikes within bursts (Figure 7C). This set 11 ABCDEFGHIJKLMNOPtonic spikingphasic spikingtonic burstingphasic burstingmixed modetype Itype IIspike latencyresonatorintegratorrebound spikerebound burstvariabilitybistability Ibistability II50 msspike frequencyadaptationthreshold Figure 7: Changes in a single component of each filter can produce a variety of behav- iors. Center: Amplitudes for a single component of the stimulus filter (ordinate) or post-spike filter (abscissa) were varied. Responses were simulated for 15 trials of a step stimulus, and the most common behavior produced is indicated by color. Small panels show example filters at each extreme of the range tested. A-D: Example responses (gray) to step stimulus (blue) for GLMs with filters indicated by corresponding letter in center panel. of behaviors could be achieved by simply sweeping over the amplitude of a single basis vector in each filter. Incorporating shifts to the basis vectors or additional basis vectors would likely be necessary to achieve more complex behaviors, such as bistability. Although we have drawn clear borders at the transition between behaviors, these transi- tions in fact occur gradually. Near the border between phasic spiking and phasic bursting, for example, there will be some trials where a single spike is produced and other trials where a burst is elicited. We have indicated the behavior that is produced most frequently here for sim- plicity. The transition from tonic bursting to tonic spiking also occurs gradually, with the near perfectly regular bursting breaking down into more irregular firing until no apparent bursts are produced. If the post-spike filter is made even more negative than the range explored in this fig- ure, the timing of tonic spiking becomes near perfectly regular as well. This is easily explained by the fact that as the post-spike filter becomes more and more negative, it imposes stronger refractoriness on the cell, which results in more regular spike timing. As the post-spike filter component amplitude is changed, there is therefore a gradual change from precisely timed bursts, to irregular firing, to precisely timed tonic spiking. In the following section, we further explore questions of spike timing precision in the GLM. 3.7 Generalization to new stimuli The behaviors shows in Figure 6 were all fit using stimuli that probe only a small range of the possible behaviors of the neuron. For example, many were probed using only a single step height. A natural question that arises is therefore: how well will these fitted GLMs generalize to predict the Izhikevich neuron responses to new stimuli? We examined this question for three canonical Izhikevich neurons from our study (Figure 8). We first generated responses from a 12 component 2 amplitudecomponent 1 amplitudeABCDABDquiescentphasic spikingphasic burstingtonicburstingtonicspiking50 ms50 ms200 msC-1-0.500.51-1.5-1-0.50 Figure 8: GLMs have lim- ited ability to genearlize to new stimuli. A: Inputs used to generate responses from Izhikevich neurons. Step amplitudes were 7, 14, & 28; 0.3, 0.6, & 1.2; and 5, 10, & 20 for B, C, & D, respectively. Standard deviation of noise was 7.2 for all neuron types. B: Top: Responses of a regu- lar spiking Izhikevich neuron to the above stimuli. Mid- dle: Spike responses of the Izhikevich neuron (red) and GLM (black) fit on responses to only the middle step size (indicated by gray box). Bot- tom: Spike responses of the Izhikevich neuron (red) and GLM (black) fit on responses to all three step sizes (indi- cated by gray box). C: Re- sponses of a phasic bursting Izhikevich neuron. All pan- els as in B. D: Responses of a tonic bursting Izhikevich neuron. All panels as in B & C. 13 020040002004000200400024602004000246-100-50050ABCD-100-50050024602004000246020040002004000200400002004000200400020040000200400time (ms)mVresponsesresponsesmVresponsesresponsesmVresponsesresponsesinput-100-50050024602004000246020040002004000200400 regular spiking Izhikevich neuron using three step heights and one noise stimulus (Figure 8A; B, top). We then simulated responses to these stimuli using the GLM fit only to the intermediate step height (Figure 8B, middle). (This is the same fit as Figure 2.) The GLM responses nearly perfectly capture the Izhikevich responses for the original stimulus, and the GLM maintains regular firing patterns for the other step heights. However, the GLM's firing rate is too high for the small step and too low for the large step. While the GLM accurately captures some firing events for the noise stimulus, the firing rate is overall too high. We next refit a GLM to responses from all three step heights, using the same set of basis vectors as the original fit (Figure 8, bottom). The responses to all three step heights are captured nearly perfectly. Additionally, the response to the noise stimulus is much more similar to that of the Izhikevich neuron. Although there is one firing event that occurs at a delay, the regular spiking GLM fit on an enriched stimulus set is better able to generalize to the noise stimulus. We performed this same test for neurons showing phasic bursting (Figure 8C) and tonic bursting (Figure 8D). (Note that although some responses of the phasic bursting Izhikevich neuron are not actually phasic, we retain the naming convention given to this set of parameters in the original paper.) For phasic bursting, the GLM fit on additional step sizes improves the accuracy of responses to the smallest and largest steps, while decreasing the accuracy of responses for the original step of intermediate size. There is marked improvement in the accuracy of responses to the noise stimulus, with many firing events being accurately captured and the GLM no longer exhibiting runaway excitation. For tonic bursting, the refit GLM retains bursting behavior but fails to even capture responses for the steps on which it was trained. As noted above, when refitting we used the same set of basis vectors as the initial fits for fair comparison. It is possible that by increasing the number of basis vectors used or tuning their properties that better fits to all stimuli might be achieved. Taken together, these results show that while GLMs might retain some characteristic re- sponse features (such as bursting) when probed with new stimuli, they often have limited ability to generalize beyond stimuli on which they are directly fit. 4 Spiking precision and reliability A noteworthy feature of the spike trains of the GLM neurons considered above is their high degree of spike timing precision and reliability across trials. This precision arises from the fact that the conditional intensity (or instantaneous spike rate) rises abruptly at spike times (due to filter outputs passing through a rapidly accelerating exponential nonlinearity), and decreases immediately after each spike due to suppressive effects of the post-spike filter. By contrast, a Poisson GLM without recurrent feedback, more commonly known as a linear-nonlinear Pois- son (LNP) cascade model, cannot produce temporally precise spike responses to a constant stimulus because its output is constrained to be a Poisson process. Real neurons, however, seem to be capable of both response modes: they emit precisely- timed spikes in some settings and highly variable spike trains in others. A seminal paper by Mainen & Sejnowski illustrated this duality by showing that spike responses to a constant DC current exhibit substantial trial-to-trial variability, whereas responses to a rapidly fluctuating injected current are precise and repeatable across trials [29]. Deterministic models like the Izhikevich model cannot, of course, mimic this property because their spikes are perfectly 14 Figure 9: Stimulus-dependent spike timing reliability. A: Top: Weak step stimulus. Bottom: Spike train responses of tonic-spiking GLM on 15 repeated trials. Although the first spike is precise and reliable, subsequent spikes have irregular timing from trial to trial. B: Top: Rapidly fluctuating stimulus. Bottom: Spike response of same GLM neuron on 15 repeated trials, exhibiting a high degree of precision and reliability. (Compare to Figure 1 of [29].) reproducible for any stimulus. (A stochastic version of the Izhikevich model with an appropriate level of injected noise could likely overcome this shortcoming, however. See [44] for a similar case in a Hodgkin-Huxley neuron.) Here we show that the GLM naturally reproduces the same form of stimulus-dependent changes in precision and reliability observed in real neurons. Figure 9 shows that a single GLM (with parameters identical to those fit to the tonic-spiking Izhikevich neuron, shown in Figure 2) produces irregular spiking in response to a constant stimulus with low-to-intermediate amplitude, and precisely-timed, reliable spikes in response to a stimulus with large, rapid fluctuations. 5 GLMs can produce super-Poisson variability We have shown that GLM neurons can reproduce the high degree of spike timing precision found in real neurons stimulated with injected currents. However, a variety of studies have reported that neurons exhibit overdispersed responses, or greater-than-Poisson spike count variability in response to repeated presentations of a sensory stimulus [47, 46, 45, 17]. A prominent recent study from Goris, Movshon, & Simoncelli showed that the degree of overdis- persion grows with mean spike count, so that the Fano factor (variance-to-mean ratio) is an increasing function of spike rate [17]. They proposed a doubly stochastic model to account for this phenomenon, in which the rate of a Poisson process is modulated by a slowly fluctuating stochastic gain variable g. For each trial, g is drawn from a gamma distribution with mean 1 and variance σ2 g. (See Methods for details.) We sought to determine if a GLM with spike-history dependence can also account for the mean-dependent overdispersion found in neural responses. To test this possibility, we simulated spike trains from the doubly stochastic model of Goris et al for three different settings 15 0369020040060080051015time (ms)constant inputfluctuating input-50050100020040060080051015time (ms)ABstimulus amplitudetrial g with the same mean spike rate (100 spikes/s), and fit a GLM of the over-dispersion factor σ2 g. We then simulated responses from each to the spike trains associated with each value of σ2 GLM to 500 ms pulses at a number of different input intensities, with each point in Figure 10 corresponding to a different intensity. We found that the GLM can indeed match the qualitative behavior of the Goris et al model, giving approximately Poisson responses at low spike rates and increasingly overdispersed re- sponses at higher rates (Figure 10C). To match the data from larger values of overdispersion g (darker curves in Figure 10C), the GLM relies on increasing amounts of self-excitation factor σ2 from the post-spike filter, but exhibits no changes in stimulus filter (Figure 10A-B). The filters here do not include an absolute refractory period, as the original model does not incorporate one. However, similar results can be achieved when a refractory period is enforced in the train- ing data. This will result in post-spike filters with strongly negative lobes on short timescales, which impose refractoriness, but otherwise similar filters to those in Figure 10. Unlike models with purely suppressive spike history effects, which capture effects due to refractoriness and reduce variability in firing (e.g., [3, 49]), here we show that allowing spike history effects to be excitatory can result in increased variability. While the former might be suitable for early sen- sory areas, such as the retina, the latter better captures the super-Poisson variability observed in higher visual areas. Intuitively, the GLM generates overdispersed spike counts because of dependencies the spike-history filter induces between early and late spikes during a trial: if the GLM neuron generates a larger-than-average number of spikes early in a trial, the positive post-spike filter produces a higher conditional intensity (and hence more spiking) later in the trial; conversely, if a neuron emits fewer-than-average spikes early in a trial, the conditional intensity will be lower later in the trial (yielding less spiking). The Goris et al model can be seen to capture similar dependencies between early and late spikes via the stochastic gain variable g, which is con- stant during a trial but independent across trials. Thus, it is reasonable to view g in the Goris et Figure 10: GLMs can produce super-Poisson variability. A: Stimulus filters for GLMs trained on three different levels of variability: high (dark blue), medium (medium), and low (light) super- Poisson variability. Spike count mean was identical in the three cases: 100 spikes/s. B: Post- spike filters for high (dark red), medium (medium), and low (light) super-Poisson variability. C: Spike count variance versus mean for three levels of variability. This relationship is strikingly similar to that observed in many cortical neurons. 16 0204060801000100200300400spike count meanPoisson spikingspike count variance-50-25000.10.2020040000.02time (ms) = -0.84 = -1.11 = -0.86time (ms)Astimulus filtersspike-history filtersBC al model as a proxy for the accumulated self-excitation from spike-history filter outputs under a GLM. We note, however, that attempts to drive the GLM to higher levels of overdispersion (e.g., Fano factors significantly > 3) often resulted in runaway self-excitation, indicating that GLMs may require additional mechanisms to maintain stability in order to produce highly over- dispersed responses through recurrent excitation alone [11, 19]. An alternative mechanism for generating over-dispersed responses with GLMs is through the addition of latent stochastic inputs [42], an avenue we have not explored here. 6 Discussion We have shown that recurrent Poisson GLMs can capture an extensive set of behaviors exhib- ited by biological neurons, including tonic and phasic spiking, bursting, spike frequency adap- tation, type I and type II behavior, and bistability. GLMs can also reproduce widely varying levels of response stochasticity, ranging from precisely timed spikes with negligible trial-to-trial variability, to substantially super-Poisson spike count variability. We have also shown that, like real neurons, GLMs can exhibit irregular firing in response to a constant stimulus, but precise and repeatable firing patterns in response to a temporally varying stimulus. Thus, generalized linear models are able to capture a rich array of spiking behaviors like many dynamical models, while remaining tractable to fit to neural data. 6.1 Relationship to previous work As mentioned above, GLMs have strong connections to a number of other models. It is partic- ularly worth noting the connection between GLMs and generalized integrate-and-fire models, such as the spike response model (SRM) extensively studied by Gerstner and colleagues [15, 13]. These models draw on much earlier work which incorporated a variable threshold that depends on spiking history [51, 10]. The SRM includes a membrane filter (analogous to the stimulus filter here) and both a spike afterpotential and moving threshold (which can be combined and are analogous to the spike history filter here). In its simplest formulation, the SRM is a deterministic model. Although the threshold for spiking can shift as a function of spike history, a spike will occur precisely at each threshold crossing. Extensions of the model have incorporated so-called "escape noise," where spiking no longer occurs deterministically at threshold crossings, but rather the probability of spiking de- pends on the distance of the membrane voltage to threshold [15, 25]. This variant of the SRM is in fact a GLM, and it is therefore worthwhile to consider how previous work investigating the SRM with escape noise relates to our results here. Early work demonstrated that such a model was capable of producing responses with high temporal precision, including both tonic spiking as well as tonic bursting [15], though demonstrating the range of behaviors that could be produced by the SRM was not the focus of this study. Additional work demonstrated that the model could produce highly repeatable spike trains to a noisy stimulus (similar to Figure 9B, though no comparison of irregular firing in response to a constant stimulus was shown) [25]. Other studies have shown that the SRM can capture the detailed statistics of neural responses [31, 41]. Further, many of these studies show that the spike responses model can be used to capture not only the relationship between an external stimulus and a neuron's response, 17 but also to faithfully capture the relationship between intracellularly recorded neural responses and injected current [25, 41]. 6.2 Limitations Despite their many advantages, GLMs have several limitations that bear further discussion. First, GLMs often do not generalize well across stimulus distributions; models fit with a partic- ular set of stimuli often do not accurately predict responses to stimuli with markedly different statistics (e.g., stimuli with large changes in mean or variance, or white noise vs. naturalistic stimuli) [18]. Secondly, GLMs often lack clear interpretability in terms of underlying mechanisms. This stands in contrast to dynamical models designed to capture specific biophysical variables and processes. In the two-dimensional Izhikevich model, for example, one variable (v) represents the neuron's membrane potential, and the other variable (u) can be understood as a membrane recovery variable, which reflects K+ channel activation and Na+ channel inactivation. Despite the fact the GLM filters do not represent specific biophysical variables, in some cases they can still provide insight into underlying biological processes. For example, recent work provides an interpretation of the GLM as a synaptic conductance based model with linear sub-threshold dynamics [27]. Work on the SRM has shown that by dividing the effects of spike history into a dynamic threshold and a spike afterpotential, one can in fact measure their separate contribu- tions with intracellular recordings of a neuron's subthreshold voltage; the spike afterpotential can be observed directly in this voltage trace, while the effects on thereshold can be estimated indirectly by noting the absence of firing [31, 41, 14]. A third known limitation of GLMs is that they lack the flexibility to capture some nonlinear response properties of real spike trains. For example, as point neuron models, GLMs do not reflect the fact that neurons often receive spatially segregated inputs on the dendritic tree, and these inputs can be processed separately and combined nonlinearly [28]. Some extensions of the GLM that incorporate nonlinear inputs and multiple subunits [8, 43, 36, 7, 30, 1, 52] may begin to address this issue, but certainly fall short of capturing the full complexity of dendritic processing. For the range of dynamical behaviors considered here, however, we did not find these extensions to be necessary. For all results shown, we used a GLM with an exponential nonlinearity. To test the depen- dence of our results on the form of the nonlinearity, we also fit GLMs to several of the behaviors with a "soft-rectifying" nonlinearity given by f (x) = log (1 + exp (x)). This function grows only linearly for large input values, but still has an exponential decay on its left tail and remains in the family of nonlinearities (convex and log-concave) for which the GLM log-likelihood is provably concave [34]. For the behaviors tested (tonic spiking, tonic bursting, phasic spiking, and phasic bursting), our results were similar to those with an exponential nonlinearity, though generally not as temporally precise. This increased precision is likely due to the fact that an exponential nonlinearity rises more steeply than a linear-rectifying function, causing the condi- tional intensity to accelerate more rapidly from a low-probability to a high-probability of spiking regime. Past studies have found that responses of both retinal ganglion cells and neocortical pyramidal neurons are well described by a GLM with exponential nonlinearity [40, 25]. It is worth noting that for many of the dynamic behaviors studied here, the GLM parameters 18 were not strongly constrained by the training data. (See "Sensitivity to changes in parameter values" in Appendix.) Slight changes in the model parameters did not produce noticeable changes in response, at least for the stereotyped range of input currents and output spike patterns considered. The filter parameters were therefore only weakly identifiable, which cor- responds to a likelihood function with a very gradual falloff along certain directions in parameter space. This uncertainty potentially complicates interpretation of the filters in terms of functions performed by the underlying biophysical mechanism. Conversely, it reveals that the suite of behaviors considered by Izhikevich and others can be achieved by a range of different GLMs, and that a richer set of input-output patterns is needed to identify a unique set of GLM param- eters. Conclusion The GLM has the ability to mimic a wide range of biophysically realistic behaviors exhibited by real neurons. Although it is clear there are some forms of nonlinear behavior it cannot produce, such as frequency-doubled responses of cat Y cells or V1 complex cells to a contrast-reversing grating, our work provides an existence proof for its ability to exhibit an important range of re- sponse types considered previously only in biophysics and applied math modeling literature. Moreover, by considering response stochasticity as another dimension along which real neu- rons vary, we have shown that that the GLM can generate response characteristics ranging from quasi-deterministic to greater-than-Poisson variability. The GLM therefore provides a flexible yet powerful tool for studying the dynamics of real neurons and the computations they carry out. Appendix MATLAB code used to generate example responses from Izhikevich neurons and to fit GLMs to these responses is available in a Github repository (https://github.com/aiweber/ GLM_and_Izhikevich). Izhikevich model simulations To generate training data for fitting the GLM, we simulated responses from an Izhikevich model [22] with parameters set to published values given for each behavior in [23] (Table 1; parameter values can be found at http://www.izhikevich.org/publications/izhikevich.m). For each behavior, we generated approximately 20 seconds of training data using the forward Euler method with fixed time step size (dt) given in Table 1. It should be noted that in some cases, published pa- rameter values did not produce the desired qualitative behavior. In these cases, we tuned the simulation parameters to achieve the desired behavior. Parameters marked with an asterisk in Table 1 indicate those that differ from published values for the corresponding behavior in [23]. Additionally, some behaviors of the Izhikevich neuron are not robust to small changes in stim- ulus timing, stimulus amplitude, or time step of integration. In particular, we found bistability to 19 be highly dependent on the precise stimulus timing (onset and duration), stimulus amplitude, and integration window. We tuned these values by hand to produce the desired behavior. GLM fitting and simulations A generalized linear model for a single neuron attributes features of a spike train to both stimu- lus dependence and spike history. Stimulus dependence is captured by a stimulus filter k, and spike-history dependence is captured by a post-spike filter h. k and h are represented with a raised cosine basis to reduce to the dimensionality necessary to fit and ensure smoothness of the filters. Basis vectors are of the form: 1 2 cos(a log[t + c] − φj) + bj(t) = (6) for t such that a log(t + c) ∈ [φj − π, φj + π] and 0 elsewhere. The parameter c determines the extent to which peaks of the basis vectors are linearly spaced, with larger values of c resulting in more linear spacing. We typically used 6 such basis vectors to fit a 100 ms stimulus filter k and 8 basis vectors to fit a 150 ms post-spike filter h, for a total of 15 parameters (including 1 2 a mixed mode type I type II spike latency neuron type tonic spiking phasic spiking tonic bursting phasic bursting b 0.02 0.2 0.02 0.25 0.02 0.2 0.02 0.25 0.2 0.02 spike frequency adaptation 0.01 0.2 -0.1 0.02 0.26 0.2 0.02 0.2 0.26 0.1* 0.02 -0.1 0.03 0.25 0.03 0.25 0.03 0.25 1.5 1.5 rebound spike rebound burst bistability I bistability II resonator integrator threshold variability 1 1 c -65 -65 -50 -55 -55 -65 -55 -65 -65 -60 -66* -60 -52 -60 -60 -60 d 6 6 2 0.05 4 5* 6 0 6 -1 6 4 0 4 0 0 I 14 0.5 10* 0.6 10 20* 25 0.5 3.49* 0.3 27.4 -5 -5 2.3 30* 40 dt (ms) 0.1 0.1 0.1 0.1 0.1 0.1 1 1 0.1 0.5 0.5 0.1 0.1 1 0.05 0.05 Table 1: Parameters of the Izhikevich neuron for dynamic behaviors shown in Figures 2-6, 8- 9, & 11. Parameters marked with * indicate parameters that differ from those used in [23]. Additionally, only a single form of bistability (bistability I) was presented in [23]. 20 In some cases, as few as 7 or as many as one for µ that determines baseline firing rate). 26 parameters were used to fit an individual Izhikevich neuron's behavior. In general, the fewest number of basis vectors required to reproduce a given behavior were used, though it is likely that by altering specific features of the basis vectors (e.g., their spacing), even fewer parameters would suffice. We fit the model parameters (weights on the basis functions for k, weights on the basis functions of h, and µ) by maximizing the log-likelihood: log λ(t) − ∆ L(θ) = (cid:88) (cid:88) λ(t) (7) t=spike t where ∆ is the time resolution of y(t). We used MATLAB's fminunc function, part of the MATLAB optimization toolbox, to find the global maximum of the likelihood function. We simulated the GLM response in time bins of the same size as the corresponding Izhike- vich neuron and computed the single-bin probability of a spike as P (y(t) ≥ 1λ(t)) = 1 − P (y(t) = 0λ(t)) = 1 − exp(∆λ(t)), (8) where ∆ is the time bin size, so that the probability of 0 or 1 spikes in a bin sums to 1 (resulting in a Bernoulli approximation to the Poisson process), disallowing spike counts greater than 1 in a single bin. Systematic variation of filter amplitudes In order to more carefully examine the transitions between different behaviors as the stimulus and post-spike filter change, we systematically varied the amplitude of individual filter compo- nents and observed the behavior produced. Each filter was parameterized with 2 components. The amplitude of one was fixed while the amplitude of the other was varied. For the stimulus filter, the amplitude of the second component was varied (-1.5 to +0.25). This allowed us to transition from monophasic to biphasic filters. The amplitude of the first component was set to be positive (+1), creating an "ON" filter appropriate for a positive step stimulus. For the post- spike filter, the amplitude of the first component was varied (-1 to +1). The amplitude of the second component was set to be negative (-3), ensuring that spiking would be suppressed on longer timescales. For the post-spike filter we also imposed an absolute refractory period of 5 ms. Finally, we included a negative baseline drive (µ = -1) to suppress spontaneous spiking so that the baseline firing rate was zero. We simulated responses to 25 identical step stimuli for each set of filters and then classi- fied the behaviors as quiescent, phasic spiking, phasic bursting, tonic spiking, or tonic bursting. The most commonly observed behavior over the 25 repetitions is depicted in Figure 7. Re- sponses were classified in the following way. If no spikes were elicited in the first 200 ms of stimulus presentation and fewer than 5 spikes were elicited during the final 10 seconds of stim- ulus presentation, the behavior was classified as quiescent. If at least one spike was elicited in the first 200 ms following stimulus onset and fewer than 5 spikes were elicited during the final 10 seconds of stimulus presentation, the behavior was classified as phasic. Phasic firing patterns were further classified into phasic spiking if only 1 or 2 spikes were elicited in the first 200 ms, and phasic bursting if 3 or more spikes were elicited in the first 200 ms. The 21 remaining responses were classified as either tonic spiking or tonic bursting in the following manner. Inter-spike interval distributions were fit with a Gaussian mixture distribution using MATLAB's gmdistribution function. We fit both a single Gaussian distribution as well as a mixture of two Gaussians and then compared the Akaike information criterion (AIC) values to determine whether the ISI distribution was better fit as a unimodal distribution or a bimodal If 0.9 · AICunimodal < AICbimodal, the spike distribution, with a lower AIC indicating better fit. train was classified as tonic spiking; otherwise, it was classified as tonic bursting. (We added the 0.9 factor to create a more stringent standard for what is classified as bursting activity so that the responses that fall into this category are strongly bimodal distributions that would be readily identified as bursting. Slightly altering the value of this factor, or eliminating it entirely, gives the same qualitative results, but merely shifts the boundary in Figure 7 between the tonic spiking and tonic bursting regions.) Figure 11: Sensitivity to changes in parameter values for a regular spiking (left) and tonic bursting (right) GLM. A: Fit coefficient of each eigenvector of Hessian matrix of likelihood, normalized by corresponding eigenvalue. Eigenvectors are in order of decreasing eigenvalues (not necessarily decreasing z-scored eigenvalues). B: Stimulus filter (top, blue) and spike history filter (bottom, red), along with two most constrained eigenvectors. These correspond to the largest (dark gray) and second largest (light gray) eigenvalues. Eigenvectors are scaled to size comparable with filters. C: Same as B, for least constrained eigenvectors. D-F: Same as A-D for tonic bursting neuron. 22 DEF246810121410ï1100101102eigenvector0255075100-0.200.20.40.6most constrained0255075100-300-200-1000100time (ms)0255075100least constrained0255075100time (ms)tonic burstingz-scoredcoefficientfilter amplitudefilter amplitudeAB246810121410ï810ï4 100 eigenvector0255075100-1012most constrained050100150-1000-5000time (ms)0255075100least constrained050100150time (ms)regular spikingCfilter amplitudez-scoredcoefficientfilter amplitude Sensitivity to changes in parameter values We wished to investigate how well constrained different features of our fit GLMs were. To do so, we calculated the eigendecomposition of the Hessian matrix of the likelihood function: H = QΛQ−1. The Hessian matrix provides a local quadratic approximation to the likelihood function, with eigenvectors qi pointing along the principal axes and length of these axes propor- tional to 1√ . Thus, larger magnitude eigenvalues indicate greater curvature (i.e., shorter axes) λi and better constrained directions, while smaller magnitude eigenvalues indicate lower curva- ture and more poorly constrained directions. Results of this analysis are shown in Figure 11 for both a regular spiking neuron (left) and tonic bursting neuron (right). For both neurons, the least constrained directions correspond to eigenvectors similar in shape to the best-fit filters or the absolute refractory period. As such, perturbations in these directions do not result in large changes in the behavior. Perturbations along eigenvectors corresponding to the most con- strained directions, on the other hand, would result in significant changes to the filter shape. The difference in scale between Panel A and Panel D indicates that overall, parameters for the tonic bursting neuron are more constrained than those for the regular spiking neuron. Doubly stochastic model with super-Poisson variability In Figure 10, we used a negative binomial model to generate spike trains with greater-than- Poisson variability [38, 17]. The negative binomial distribution can be conceived as a doubly- stochastic model in which the rate of a Poisson process is modulated by an iid gamma random variable on each trial. Following [17], we modeled responses with a stochastic gain variable g with mean 1 and variance σ2 g that obeys a gamma distribution: P (gr, s) = gr−1 exp 1 srΓ(r) (cid:17) (cid:16)−g s , (9) where s = σ2 represents the gamma function. g denotes the scale parameter, r = 1/σ2 g is the shape parameter, and Γ(·) The spike count conditioned on g and a stimulus S for each trial then obeys a Poisson distribution: P (yg, S) = (∆gf (S))y y! exp (−∆gf (S)) , (10) where ∆ is the time bin size, g is the gain, and f (S) is the tuning curve that specifies the mean response to stimulus S. g = 0, the gain g is deterministically equal to 1 and the spike count is Poisson with mean and variance equal to ∆f (S). For responses with g > 0, however, responses are overdispersed relative to the Poisson distribution and have σ2 mean ∆f (S) and variance ∆f (S)(1 + σ2 In the limit σ2 For the results shown in Figure 10, we simulated data from the negative binomial dis- g ∈ tribution with a single mean rate f (S) (100 spikes/s) at three different gain variances σ2 {.0125, 0.02, 0.05}. Spike counts were drawn iid across trials, with spike times distributed uni- formly within each trial to generate spike trains suitable for GLM fitting. We used these spike trains to fit a GLM to the data associated with each value of σ2 g, with an assumed constant input current for each trial. g∆f (S)). 23 Acknowledgments We would like to thank Adrienne Fairhall, Sara Solla, and James Fitzgerald for helpful com- ments and discussions. References [1] Ahrens, M. B., Linden, J. F., and Sahani, M. (2008). Nonlinearities and contextual influ- ences in auditory cortical responses modeled with multilinear spectrotemporal methods. J Neurosci, 28(8):1929–1942. [2] Babadi, B., Casti, A., Xiao, Y., Kaplan, E., and Paninski, L. (2010). A generalized linear model of the impact of direct and indirect inputs to the lateral geniculate nucleus. Journal of Vision, 10(10):22. [3] Berry, M. and Meister, M. (1998). Refractoriness and neural precision. Journal of Neuro- science, 18:2200–2211. [4] Brette, R. and Gerstner, W. (2005). Adaptive exponential integrate-and-fire model as an effective description of neuronal activity. Journal of neurophysiology, 94(5):3637–3642. [5] Calabrese, A., Schumacher, J. W., Schneider, D. M., Paninski, L., and Woolley, S. M. N. (2011). A generalized linear model for estimating spectrotemporal receptive fields from responses to natural sounds. PLoS One, 6(1):e16104. [6] Cox, D. and Isham, V. (1980). Point Processes. Chapman & Hall/CRC Monographs on Statistics & Applied Probability. Taylor & Francis. [7] Cui, Y., Liu, L. D., Khawaja, F. A., Pack, C. C., and Butts, D. A. (2013). Diverse sup- pressive influences in area mt and selectivity to complex motion features. The Journal of Neuroscience, 33(42):16715–16728. [8] Fitzgerald, J. D., Rowekamp, R. J., Sincich, L. C., and Sharpee, T. O. (2011). Second order dimensionality reduction using minimum and maximum mutual information models. PLoS Comput Biol, 7(10):e1002249. [9] FitzHugh, R. (1961). Impulses and physiological states in theoretical models of nerve membrane. Biophysical journal, 1(6):445. [10] Geisler, C. D. and Goldberg, J. M. (1966). A Stochastic Model of the Repetitive Activity of Neurons. Biophysical Journal, 6(1):53–69. [11] Gerhard, F., Deger, M., and Truccolo, W. (2017). On the stability and dynamics of stochas- tic spiking neuron models: Nonlinear hawkes process and point process glms. PLoS com- putational biology, 13(2):e1005390. [12] Gerstner, W. (1995). Time structure of the activity in neural network models. Physical Review E, 51(1):738–758. 24 [13] Gerstner, W. (2001). A framework for spiking neuron models: The spike response model. In Moss, F. and Gielen, S., editors, The Handbook of Biological Physics, volume 4, pages 469–516. [14] Gerstner, W., Kistler, W. M., Naud, R., and Paninski, L. (2014). Neuronal Dynamics: From Single Neurons to Networks and Models of Cognition. Cambridge University Press, New York, NY, USA. [15] Gerstner, W. and van Hemmen, J. (1992). Associative memory in a network of 'spiking' neurons. Network: Computation in Neural Systems, 3(2):139–164. [16] Gerstner, W., van Hemmen, J. L., and Cowan, J. D. (1996). What matters in neuronal locking? Neural computation, 8:1653–1676. [17] Goris, R. L. T., Movshon, J. A., and Simoncelli, E. P. (2014). Partitioning neuronal vari- ability. Nat Neurosci, 17(6):858–865. [18] Heitman, A., Greschner, M., Field, G., Li, P., Ahn, D., Sher, A., Litke, A., and Chichilnisky, In E. (2014). Representation and reconstruction of natural scenes in the primate retina. Computational and Systems Neuroscience (CoSyNe) Abstracts, pages 134 – 135. [19] Hocker, D. and Park, I. M. (2017). Multistep inference for generalized linear spiking mod- els curbs runaway excitation. In 8th International IEEE EMBS Conference On Neural Engi- neering. [20] Hodgkin, A. (1948). The local electric changes associated with repetitive action in a non- medullated axon. The Journal of physiology, 107(2):165–181. [21] Hodgkin, A. L. and Huxley, A. F. (1952). A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol., 117(4):500–544. [22] Izhikevich, E. M. (2003). Simple model of spiking neurons. IEEE Trans Neural Netw, 14(6):1569–1572. [23] Izhikevich, E. M. (2004). Which model to use for cortical spiking neurons? IEEE Trans Neural Netw, 15(5):1063–1070. [24] Jolivet, R., Lewis, T., and Gerstner, W. (2003). The spike response model: a framework to predict neuronal spike trains. Springer Lecture notes in computer science, 2714:846–853. [25] Jolivet, R., Rauch, A., Lüscher, H. R., and Gerstner, W. (2006). Predicting spike timing of neocortical pyramidal neurons by simple threshold models. Journal of Computational Neuroscience, 21(1):35–49. [26] Keat, J., Reinagel, P., Reid, R., and Meister, M. (2001). Predicting every spike: a model for the responses of visual neurons. Neuron, 30:803–817. 25 [27] Latimer, K. W., Chichilnisky, E. J., Rieke, F., and Pillow, J. W. (2014). Inferring synap- tic conductances from spike trains with a biophysically inspired point process model. In Ghahramani, Z., Welling, M., Cortes, C., Lawrence, N., and Weinberger, K., editors, Ad- vances in Neural Information Processing Systems 27, pages 954–962. Curran Associates, Inc. [28] London, M. and Häusser, M. (2005). Dendritic Computation. Annual Review of Neuro- science, 28(1):503–532. [29] Mainen, Z. and Sejnowski, T. (1995). Reliability of spike timing in neocortical neurons. Science, 268:1503–1506. [30] McFarland, J. M., Cui, Y., and Butts, D. A. (2013). Inferring nonlinear neuronal computa- tion based on physiologically plausible inputs. PLoS Comput Biol, 9(7):e1003143+. [31] Mensi, S., Naud, R., Pozzorini, C., Avermann, M., Petersen, C. C. H., and Gerstner, W. (2012). Parameter extraction and classification of three cortical neuron types reveals two distinct adaptation mechanisms. Journal of Neurophysiology, 107:1756–1775. [32] Morris, C. and Lecar, H. (1981). Voltage oscillations in the barnacle giant muscle fiber. Biophysical journal, 35(July):193–213. [33] Nagumo, J., Arimoto, S., and Yoshizawa, S. (1962). An Active Pulse Transmission Line Simulating Nerve Axon. Proceedings of the IRE, 117(m V):2061–2070. [34] Paninski, L. (2004). Maximum likelihood estimation of cascade point-process neural en- coding models. Network: Computation in Neural Systems, 15:243–262. [35] Paninski, L., Pillow, J. W., and Simoncelli, E. P. (2004). Maximum likelihood estimation of a stochastic integrate-and-fire neural model. Neural Computation, 16:2533–2561. [36] Park, I. M., Archer, E. W., Priebe, N., and Pillow, J. W. (2013). Spectral methods for neural In Advances in Neural Information characterization using generalized quadratic models. Processing Systems 26, pages 2454–2462. [37] Perkel, D. H., Gerstein, G. L., and Moore, G. P. (1967). Neuronal spike trains and stochas- tic point processes i. the single spike train. Biophysical Journal, 7(4):391–418. [38] Pillow, J. and Scott, J. (2012). Fully bayesian inference for neural models with negative- In Bartlett, P., Pereira, F., Burges, C., Bottou, L., and Weinberger, K., binomial spiking. editors, Advances in Neural Information Processing Systems 25, pages 1907–1915. [39] Pillow, J. W., Paninski, L., Uzzell, V. J., Simoncelli, E. P., and Chichilnisky, E. J. (2005). Prediction and decoding of retinal ganglion cell responses with a probabilistic spiking model. The Journal of Neuroscience, 25:11003–11013. [40] Pillow, J. W., Shlens, J., Paninski, L., Sher, A., Litke, A. M., and Chichilnisky, E. J. Simon- celli, E. P. (2008). Spatio-temporal correlations and visual signaling in a complete neuronal population. Nature, 454:995–999. 26 [41] Pozzorini, C., Naud, R., Mensi, S., and Gerstner, W. (2013). Temporal whitening by power-law adaptation in neocortical neurons. Nature Neuroscience, 16(7):942–948. [42] Rabinowitz, N. C., Goris, R. L., Cohen, M., and Simoncelli, E. (2015). Attention stabilizes the shared gain of v4 populations. eLife. [43] Rajan, K., Marre, O., and Tkacik, G. (2013). Learning quadratic receptive fields from neural responses to natural stimuli. Neural Computation, 25(7):1661–1692. [44] Schneidman, E., Freedman, B., and Segev, I. (1998). Ion channel stochasticity may be critical in determining the reliability and precision of spike timing. Neural computation, 10(7):1679–703. [45] Shadlen, M. and Newsome, W. (1998). The variable discharge of cortical neurons: im- plications for connectivity, computation, and information coding. Journal of Neuroscience, 18:3870–3896. [46] Tolhurst, D. J., Movshon, J. A., and Dean, A. F. (1983). The statistical reliability of signals in single neurons in cat and monkey visual cortex. Vision Res, 23(8):775–785. [47] Tomko, G. J. and Crapper, D. R. (1974). Neuronal variability: non-stationary responses to identical visual stimuli. Brain research, 79(3):405–418. [48] Truccolo, W., Eden, U. T., Fellows, M. R., Donoghue, J. P., and Brown, E. N. (2005). A point process framework for relating neural spiking activity to spiking history, neural ensem- ble and extrinsic covariate effects. J. Neurophysiol, 93(2):1074–1089. [49] Uzzell, V. J. (2004). Precision of Spike Trains in Primate Retinal Ganglion Cells. Journal of Neurophysiology, 92(2):780–789. [50] Weber, F., Machens, C. K., and Borst, A. (2012). Disentangling the functional conse- quences of the connectivity between optic-flow processing neurons. Nature neuroscience, 15(3):441–448. [51] Weiss, T. F. (1966). A model of the peripheral auditory system. Kybernetik, 3(4):153–175. [52] Williamson, R. S., Sahani, M., and Pillow, J. W. (2015). The equivalence of information- theoretic and likelihood-based methods for neural dimensionality reduction. PLoS Comput Biol, 11(4):e1004141. [53] Zhao, M. and Iyengar, S. (2010). Nonconvergence in logistic and poisson models for neural spiking. Neural Computation, 22(5):1231–1244. PMID: 20100077. 27
1008.1410
1
1008
2010-08-08T15:13:12
Neuronal Response Clamp
[ "q-bio.NC" ]
Since the first recordings made of evoked action potentials it has become apparent that the responses of individual neurons to ongoing physiologically relevant input, are highly variable. This variability is manifested in non-stationary behavior of practically every observable neuronal response feature. Here we introduce the Neuronal Response Clamp, a closed-loop technique enabling full control over two important single neuron activity variables: response probability and stimulus-spike latency. The technique is applicable over extended durations (up to several hours), and is effective even on the background of ongoing neuronal network activity. The Response Clamp technique is a powerful tool, extending the voltage-clamp and dynamic-clamp approaches to the neuron's functional level, namely - its spiking behavior.
q-bio.NC
q-bio
Neuronal Response Clamp immediate November 20, 2018 Avner Wallach1,2, Danny Eytan1,3, Asaf Gal1,4, Christoph Zrenner1, Ron Meir1,2 and Shimon Marom1 1 Network Biology Research Laboratories, Lorry Lokey Interdisciplinary Center for Life Sciences and Engineering, Technion, Haifa, Israel 2 Faculty of Electrical Engineering, Technion, Haifa, Israel 3 Rambam Medical Center, Haifa, Israel 4 The Interdisciplinary Center for Neural Computation (ICNC), The Hebrew Uni- versity, Jerusalem, Israel 0 1 0 2 g u A 8 ] . C N o i b - q [ 1 v 0 1 4 1 . 8 0 0 1 : v i X r a 1 Abstract Since the first recordings made of evoked action potentials it has be- come apparent that the responses of individual neurons to ongoing physiologically relevant input, are highly variable. This variability is manifested in non-stationary behavior of practically every observable neuronal response feature. Here we introduce the Neuronal Response Clamp, a closed-loop technique enabling full control over two important single neuron activity variables: response probability and stimulus- spike latency. The technique is applicable over extended durations (up to several hours), and is effective even on the background of ongoing neuronal network activity. The Response Clamp technique is a power- ful tool, extending the voltage-clamp and dynamic-clamp approaches to the neuron's functional level, namely - its spiking behavior. Keywords: Excitability ; Neuron ; Clamp 2 Introduction The responses of single neurons and neuronal populations to ongoing stimuli within a physiological range of frequencies are highly variable; this variability is manifested in both the ability to evoke an action potential and the latency between stimulus and response. Response variability was already obvious a century ago, when Adrian and Zot- terman described a "lack of regularity" in the "all-or-none" impulse response in their classical paper (Adrian & Zotterman, 1926), the first documentation of recorded action potentials. Nowadays, response vari- ability is generally accepted to be an important phenomenon (Stein, 1965; Mainen & Sejnowski, 1995; Arieli, Sterkin, Grinvald, & Aertsen, 1996; Reich, Victor, Knight, Ozaki, & Kaplan, 1997; Arsiero, Luscher, Lundstrom, & Giugliano, 2007; Soteropoulos & Baker, 2009) that re- flects the immensity of involved cellular level machineries, covering any observable cell-physiology timescale (Marom, 2010). Analyses of the sources of response variability, as well as its impact on neuronal func- tionality, is practically intractable due to the non-linearity involved in the reciprocal relations between response variability and its underlying sources. It would be highly desirable to develop a method for controlling the reliability of neuronal responses to input. By studying the sig- nals required to achieve this control one may hopefully shed light on the input-output relationships of the neuron as well as lead to a bet- ter understaning of the underlying sources of response variability. In this context, analogy to the voltage-clamp method (Hodgkin, Huxley, & Katz, 1952) immediately comes to mind. There, Hodgkin, Hux- ley and Katz studied the dependency of conductances on voltage by "clamping" the membrane potential and analyzing the required control signals (currents). By doing so, they broke the circular relationships between membrane voltage and voltage-dependent conductances; in- deed, one cannot imagine neuronal physiology without the insights that the voltage-clamp method provided throughout the years. 3 In this study we present and demonstrate the concept of Neuronal Response Clamp. The idea is to "clamp" a defined feature of the neu- ronal spiking response. This is achieved by implementing a control cir- cuit that measures pre-defined neuronal response characteristics, com- pares them to desired values, and corrects errors by changing stimu- lation features. Specifically, we devised a closed-loop real-time system that implements what is known as a Proportional-Integral-Derivative (PID) controller in system engineering (Levine et al., 1996), to com- pute the "error" signal between desired and actual neuronal responses, and correct the error by manipulating the stimulation frequency. Here we show that this system fully restrains the variability of neuronal spik- ing probability, and provides complete control over the precise latency between stimulus and response, both in synaptically isolated neurons, and within the context of active networks. 4 Materials and Methods Cell preparation Cortical neurons were obtained from newborn rats (Sprague-Dawley) within 24 hours after birth using mechanical and enzymatic procedures described in earlier studies (Marom & Shahaf, 2002). The neurons were plated directly onto substrate-integrated multi electrode arrays and allowed to develop functionally and structurally mature networks over a time period of 2-3 weeks. The number of neurons in a typi- cal network is on the order of 10,000. The preparations were bathed in MEM supplemented with heat-inactivated horse serum (5%), glu- tamine (0.5 mM), glucose (20 mM), and gentamycin (10 g/ml), and maintained in an atmosphere of 37◦C, 5% CO2 and 95% air in an incu- bator as well as during the recording phases. An array of Ti/Au/TiN extracellular electrodes, 30µm in diameter, and spaced either 500 µm or 200 µm from each other (MultiChannelSystems, Reutlingen, Ger- many) were used. The insulation layer (silicon nitride) was pre-treated with polyethyleneimine (Sigma, 0.01% in 0.1M Borate buffer solu- tion). To completely block synaptic transmission in the network, 20 µM APV (amino-5-phosphonovaleric acid), 10 µM CNQX (6-cyano-7- nitroquinoxaline-2,3-dione), and 5 µM Bicuculline were added to the bathing solution. Measurements and Stimulation A commercial amplifier (MEA-1060-inv-BC, MCS, Reutlingen, Ger- many) with frequency limits of 150-3,000Hz and a gain of x1024 was used. Rectangular 200 µs biphasic 20-50 µA current or 600-800 mV stimulation through extracellular electrodes was performed using a dedicated stimulus generator (MCS, Reutlingen, Germany). Data was digitized using data acquisition board (PD2-MF-64-3M/12H, UEI, Walpole, MA, USA). Each channel was sampled at a frequency of 16 Ksample/s (for spike probability experiments) or 96 Ksamples/s (for spike latency 5 experiments). Action potentials were detected online by threshold crossing. All spike times and shapes, as well as 15 ms voltage traces from all electrodes after each stimulus, were recorded for analyses. Data processing and closed-loop stimulation was performed using a Simulink (The Mathworks, Natick, MA, USA) based xPC target ap- plication. Computing spike probability on-line Let sn be an indicator function, so that sn = 1 if the neuron generated a spike on the nth stimulus and sn = 0 otherwise. We define π(t) as the probability of the neuron to emit a spike if it is stimulated at time t. We can estimate this probability using all past responses {si}n i=1 at times {ti}n i=1, by integrating them with an exponential kernel, (cid:101)Pn = e− tn−t0 τ + n(cid:88) i=1 si(1 − e− ti−ti−1 τ )e− tn−ti τ , where τ is the kernel's time-constant (10-30 sec in the experiments presented). To compute this on-line, we used the recursive formula: (cid:101)Pn = sn(1 − e− tn−tn−1 τ ) + (cid:101)Pn−1e− tn−tn−1 τ . PID Controller A Proportional-Integral-Derivative (PID) controller was realized on the xPC target system. The input to the controller is the error signal, en = R∗ n − Rn where R∗ n and Rn are the desired and actual response to the nth stim- ulus, respectively. The output of the controller is generally composed of three components, yn = gP en + gI ei + gD(en − en−1) n(cid:88) i=0 6 where gP , gI and gD are the proportional, integral and derivative gains, respectively (in the experiments presented in this work, gP was 25- 30Hz, gI was 0.125-0.56Hz and gD was 0-125Hz). Finally, the instan- taneous stimulation frequency (i.e. the reciprocal of the ISI) consisted of the controller's output and some 'baseline' frequency: fn = yn + fbaseline, where fbaseline was 2Hz in latency experiments and 6.67Hz in proba- bility experiments. In our system, the maximal stimulation rate was 40 Hz. In the case of failed spikes when latency was controlled, the controller maintained the same stimulation frequency that produced a spike at a previous stimulus. 7 Results The experiments described in this work were conducted in networks of rat cortical neurons, plated in-vitro on a substrate that caters to ex- tracellular recording and stimulation of electrical activity, at the single cell level. Evoked spiking activity of a given individual neuron in such a network, reflects both the direct effect of external stimulation, as well as the effects of synaptic activations by other neurons (Marom & Shahaf, 2002). We first examine the capacity of the Neuronal Response Clamp technique to control the variability of neuronal responses of neu- rons that are isolated from the network by pharmacologically blocking all major types of synaptic inputs (APV [amino-5-phosphonovaleric acid; 20 µM], CNQX [6-cyano-7-nitroquinoxaline-2,3-dione; 10 µM], and Bicuculline [5 µM]). Under complete blockade of synaptic transmission, all spontaneous activity dies out and the network becomes quiescent (Marom & Shahaf, 2002). At this state, applying a single, short (sub-millisecond) pulse between two electrodes, generates a single action potential in a subset of neurons. The dynamics of the single neuron response may thus be in- terrogated using a long sequence of stimuli. We refer to this stimulation scenario as an open loop protocol, since the (predefined) stimulation sequence is not affected by the evoked neuronal responses. In the open loop protocol, the nature of the evoked responses leads to a distinction between two qualitatively different stimulation regimes: When stim- ulation rate is low (below 5-7 Hz), response probability is practically one, i.e. the neuron responds almost to each and every stimulus, and the time-delay (latency) between the stimulus and the evoked spike is quite stable (Figure 1 middle panel). At higher stimulation rates, both response probability and response latency gradually become un- stable, exhibiting considerable, seemingly erratic fluctuations (Figure 1 bottom). Figure 2 depicts the dynamics of the two features (referred to collectively as the neural response) at three stimulation rates (1, 5, and 20 Hz) over extended timescales. Note the slow drift of response 8 latency in the 5 Hz stimulation experiment, probably reflecting slow modulations of membrane excitability. Thus, two features of neural activity, response latency and response probability, emerge as natural candidates for stabilization using the Re- sponse Clamp technique. These two features are markers of the neu- ronal excitability status, reflecting the balance between exciting and restoring membrane conductances. In order to clamp these two fea- tures, continuous, online estimation of their values is required. Unlike response latency, which can be measured individually for each of the evoked action-potentials, response probability must be estimated based on recent spiking history (i.e. using several responses). We calculate this latter measure by convolving the spiking history with an expo- nential kernel, which means that recent responses are given a greater weight (explained in Materials and Methods). The two response features mentioned above possess a useful qual- ity, from the point of view of control: they exhibit a simple, monotonic relationship with the instantaneous stimulation rate. Increasing the rate (i.e. by shortening the inter-stimulus-interval) leads to predictable outcomes, namely a decrease in spike probability and an increase in spike latency. Therefore, negative feedback may be used in order to produce a desired response, either constant or time-varying. This is achieved in the following manner (Figure 3): A response feature to be clamped is chosen (either response probability or response latency). Following each stimulus, a clamp error is computed by comparing the evoked response with its desired value. This error signal is then fed into a PID controller which accordingly alters the stimulation rate in order to decrease the error towards zero. The PID controller consists of three components: the proportional component (P) supplies the neces- sary negative feedback; the integrated component (I) counteracts slow changes in the system's responsiveness; and the derivative component (D) contributes dissipation needed to damp oscillations. For details on controller construction and implementation see Methods. We first compared the neuronal response in open-loop, where stim- 9 ulation rate is maintained constant, to the neuronal response under closed-loop conditions in synaptically isolated neurons for both re- sponse features, over a wide range of stimulation rates and response levels. In contrast to the open loop response, the response under closed- loop conditions quickly converges to the desired value (Figure 4); the clamp may be maintained for up to several hours (see below). Thus, Response Clamp counteracts the slow trends observed in the spike re- sponse latency, and restrains the huge fluctuations of the spike response probability. The response clamp technique is reliable and robust: Fig- ure 5 summarizes the results of 131 experimental blocks, showing that the standard deviation in response probability is markedly decreased under closed loop, compared to open loop stimulation regimes. The robustness of the technique is reflected in the fact that no fine-tuning of the PID gain parameters was required in order to achieve successful clamp. We were interested to determine whether it is the specific pattern of stimulation, evoked by the controller is sufficient, in and by itself to restrain response variability. We therefore played-back a stimulation pattern produced by the controller under Response Clamp conditions, to stimulate the same neuron again in an open-loop manner. Although the "replayed" input sequences were identical, stability quickly deteri- orated and the response probability diverged from the desired value (Figure 6). Thus, feedback is essential in order to achieve response stabilization: the exact temporal pattern of the stimulation series does not, in and by itself, cause stable neuronal response; rather, the instan- taneous state of the neuron must be continuously monitored and taken into account while computing the control signal. Note that the result- ing response rate (i.e. the product of stimulation rate and response probability), while not constant in either the open loop or under Re- sponse Clamp, is significantly more stable in the latter condition. Also, note the duration of the experiment presented in Figure 6, extending over 3 hours. The control signal in this long experiment, as well as in similarly long experiments (results not shown), exhibits complex 10 dynamics that expose the myriad underlying adaptive processes. In all the above mentioned experiments the response was clamped to a constant value. However, the neuronal response can also be clamped to a desired time-varying pattern. To demonstrate this capacity, we performed an experiment where the desired response latency is grad- ually ramped up and down to three different values (Figure 7). We observed that in the time-varying clamp, the rate at which the de- sired response may be changed is limited by the maximal and minimal stimulation rates; if the controller surpasses these values, its output saturates and the stimulation loop is effectively broken. In all the experiments described so far, neurons were pharmacolog- ically isolated from the network input. It was not at all obvious that the ability of the response clamp to restrain response variability would be conserved once ongoing input from the network is allowed. We thus repeated the experiments described above without blocking synaptic transmission, i.e. when the neuron is embedded in an active network. We found that the Response Clamp technique was equally effective under these conditions (Figure 8). Moreover, analyzing spikes simulta- neously recorded from many different neurons in the network (beyond the controlled neuron), revealed that the responses of these other neu- rons fall into two broad categories: Some neurons were influenced by the dynamic control over the target (controlled) neuron, whereas oth- ers seem to be completely unaffected. These results suggest that the Response Clamp technique may be used as a tool to study single neu- ron dynamics even when isolation of the neuron from its context is impossible or undesirable. 11 Discussion In this study we show that evoked neuronal spiking patterns may be controlled using a simple feedback system. The Response Clamp design was applied to control either the time delay between stimulus and an evoked spike, or the probability of evoked spikes. Control of these response variables was shown to be applicable both when the target neuron is synaptically isolated from the rest of the network, and on the background of ongoing synaptic input. Historically, the concept of feedback control proved effective in the analyses of excitability. Its use, however, was largely focused on the study of specific membrane conductances under voltage-clamp and patch-clamp conditions (Hodgkin et al., 1952; Neher, Sakmann, & Steinbach, 1978). The approach was further advanced by the dynamic- clamp experimental design to unfold the impacts of specific conduc- tances on the dynamics of the overall excitability (Sharp, O'Neil, Ab- bott, & Marder, 1993). In that context, the methodology presented here is a natural step up in the ladder of organization levels. It enables control of neuronal response patterns at the macroscopic level, without monitoring underlying microscopic variables. One should bear in mind that in the original voltage-clamp stud- ies of action potential generation the variable under control (namely, membrane potential) determines the reaction rates of all the relevant processes, so that the closed loop behavior becomes linear and time in- variant. In contrast, in the dynamic clamp method only a few compo- nents of the system are being controlled, with the hope of illuminating the role of these components in the overall behavior. In this context the method suggested here resembles the dynamic clamp in the sense that only the processes that depend on the responsiveness of the sys- tem are being clamped. Hopefully, this may aid in the identification of these processes and the evaluation of their contribution to the response dynamics. Two technical difficulties in the design of the Response Clamp must 12 be noted. First, unlike the membrane potential which may be moni- tored continuously, the neuron must be externally perturbed in order to estimate its state (i.e. responsiveness to stimulation). Therefore, the controller is always "one step behind" the system being controlled. Since these perturbations themselves affect the neuron's responsive- ness, a classic "observer effect" arises. This problem becomes more pronounced in the case of neural response probability, where the state is estimated using past spiking history. Future improvements of the state estimation process might incorporate compensation for the effects of external stimulation. A second limitation in implementing the Response Clamp technique relates to the response features to be controlled. The choice of response latency and response probability stems from the need for a predictable reaction to changes in stimulation rate. This does not necessarily re- quire a monotonic relationship; a successful clamp may be realized as long as the short term outcomes of modifying the stimulation are predictable. The Response Clamp technique offers the potential of facilitating full characterization of the input-output relationships of the neuron. This challenge is practically intractable in open loop due to the nonlin- earity of the system and the cumulative effect of underlying processes spanning a wide range of timescales. By manipulating well-defined features of the neuronal responsiveness, one may hope to control state- dependent dynamics, thus enabling the identification of the role of such processes in the overall behavior of the system. 13 Acknowledgments The authors thank Erez Braun, Naama Brenner, Danni Dagan, Steve Goldstein, Effraim Wallach and Noam Ziv for their useful comments and suggestions. 14 References Adrian, E. D., & Zotterman, Y. (1926). The impulses produced by sensory nerve endings: Part 3. impulses set up by touch and pressure. J Physiol, 61 (4), 465 -- 483. Arieli, A., Sterkin, A., Grinvald, A., & Aertsen, A. (1996). Dynamics of ongoing activity: explanation of the large variability in evoked cortical responses. Science, 273 (5283), 1868. Arsiero, M., Luscher, H., Lundstrom, B., & Giugliano, M. (2007). The impact of input fluctuations on the frequency-current relationships of layer 5 pyramidal neurons in the rat medial prefrontal cortex. Journal of Neuroscience, 27 (12), 3274. Hodgkin, A., Huxley, A., & Katz, B.(1952). Measurement of current-voltage rela- tions in the membrane of the giant axon of Loligo. The Journal of physiology, 116 (4), 424. Levine, W., et al.(1996). The Control Handbook. CRC press Boca Raton, FL. Mainen, Z., & Sejnowski, T. (1995). Reliability of spike timing in neocortical neurons. Science, 268 (5216), 1503. Marom, S. (2010). Neural timescales or lack thereof. Progress in Neurobiology, 90 (1), 16 - 28. Marom, S., & Shahaf, G. (2002). Development, learning and memory in large lessons beyond anatomy. Quarterly random networks of cortical neurons: Reviews of Biophysics, 35 (01), 63 -- 87. Neher, E., Sakmann, B., & Steinbach, J.(1978). The extracellular patch clamp: a method for resolving currents through individual open channels in biological membranes. Pflugers Archiv European Journal of Physiology, 375 (2), 219 -- 228. Reich, D., Victor, J., Knight, B., Ozaki, T., & Kaplan, E. (1997). Response variability and timing precision of neuronal spike trains in vivo. Journal of neurophysiology, 77 (5), 2836. Sharp, A., O'Neil, M., Abbott, L., & Marder, E. (1993). Dynamic clamp: computer-generated conductances in real neurons. Journal of neurophysi- ology, 69 (3), 992. Soteropoulos, D., & Baker, S.(2009). Quantifying Neural Coding of Event Timing. Journal of Neurophysiology, 101 (1), 402. Stein, R.(1965). A theoretical analysis of neuronal variability. Biophysical Journal, 5 (2), 173 -- 194. 15 Figures 16 Figure 1: Evoked responses of an isolated neuron in-vitro to periodic stimulation in open loop. The top panel (a) shows a single extracellular voltage trace recorded for 20 ms, beginning at the onset of stimulation. The stimulus artifact lasts for 1-2 ms (shaded area). Response latency is the time elapsed from the onset of stimulation to the detected peak of the action potential; in our system, response latency is in the range of 3-10 ms. (b) Traces (every 10th, shown top to bottom) in response to a low (1 Hz) stimulation rate. Spiking under this condition is highly regular and the latency is constant. (c) Traces (every 10th, shown top to bottom) of the same neuron when stimulation rate is high (20 Hz). Both response probability and latency rapidly become irregular under this condition. 17 Figure 2: Response latency (a) and response probability (b) for three open loop stimulation rates. At extremely low stimulation rate (1 Hz, depicted yellow), response is highly reliable (i.e. response probability is one) and response latency is practically constant. At higher stimulation rate (5 Hz, depicted purple), response probability is also one, but slow trends in the response latency appear, probably reflecting slow processes of inactivation. At high stimulation rate (20 Hz, depicted blue) both response probability and response latency are highly unstable. The response probability was computed using 2 s bins. 18 Figure 3: General scheme of a PID controller, designed for clamping neuronal re- sponses (stimulus-spike time delay, or response probability). The error signal is calculated (feedback), and subjected to three different transformations that addi- tively dictate the nature of stimulation needed for clamping the response. This is standard control algorithm was implemented within a Simulink environment. 19 Figure 4: Demonstration of the neuronal Response Clamp. In each of the panels, two experiments are shown (depicted by two different colors). Top row shows examples of controlling response probability (a) and stimulus-spike time delay (b). Two clamped values are demonstrated in each of these panels. Note the stability of the clamped response. In contrast, when the average stimulation rates used for the clamping procedure of the two top panels are applied under open loop conditions (c and d), the response develops marked fluctuations. This is especially apparent when response probability (c) is considered. 20 Figure 5: The Response Clamp consistently decreases the variability of response probability. Neuronal response probability was clamped to different values ranging from 1 to 0.25, for 10 to 15 minutes. The response's standard deviation during this period is compared with that of the same neuron, when the mean rate used during the clamp period is re-applied in an open loop design (see the examples depicted in Figure 4). Data from 131 such experimental sessions conducted on 3 different neurons (each in a different culture, depicted by different colors) are presented. The data are normalized to the maximal standard deviation acquired for each neuron. Histograms of the data are aligned to the relevant axes. 21 Figure 6: Demonstration of the importance of feedback in the stabilization of neuronal response probability. Stimulation rate (a), response probability (b) and response rate (c) at two stimulation scenarios are depicted. First, the neuron's response probability was clamped to a constant value (0.67) for three hours (blue). Then, the stimulation pattern generated by the controller was replayed in open loop (purple). Stability of both response probability and response rate is obtained only in the presence of feedback. 22 Figure 7: Demonstration of time varying Response Clamp. Response latency was clamped, while the desired value (depicted blue) was gradually increased and decreased alternatingly. Latency is normalized so that the baseline level (i.e. the latency at an extremely low stimulation rate, in the example presented here it is 6 ms) is 0 and maximal latency detected (10 ms in this example) is set to 1. Note that at extremly high latency values the behavior becomes irregular and the clamp is effectively lost. 23 Figure 8: Time varying Response Clamp of a neuron with background synaptic input. (a) A neuron that responds (blue) directly to external stimulation was controlled to follow a sine-wave response probability pattern (12 minutes period, range between 0.75 and 0.25, depicted light green). The figure shows the response probability of another, directly stimulated yet uncontrolled neuron (purple). The response of such neurons, while not follwing the desired response closely like the controlled neuron does, is modulated by the clamp. The figure also presents the response probability of yet another neuron, which is not affected directly by the stimulation, but is activated by the synaptic inputs it recieves from the network (yellow). The activity of such neurons seems to be unaffected by the clamp. This is also apparent by looking at the spike rate histogram of 20 other neurons in the network (b, 10 s bins), demonstrating the presence of uninterrupted on-going background activity in the network. 24
1803.01236
1
1803
2018-03-03T20:45:33
Vocal effort modulates the motor planning of short speech structures
[ "q-bio.NC", "physics.bio-ph" ]
Speech requires programming the sequence of vocal gestures that produce the sounds of words. Here we explored the timing of this program by asking our participants to pronounce, as quickly as possible, a sequence of consonant-consonant-vowel (CCV) structures appearing on screen. We measured the delay between visual presentation and voice onset. In the case of plosive consonants, produced by sharp and well defined movements of the vocal tract, we found that delays are positively correlated with the duration of the transition between consonants. We then used a battery of statistical tests and mathematical vocal models to show that delays reflect the motor planning of CCVs and transitions are proxy indicators of the vocal effort needed to produce them. These results support that the effort required to produce the sequence of movements of a vocal gesture modulates the onset of the motor plan.
q-bio.NC
q-bio
Vocal effort modulates the motor planning of short speech structures Alan Taitz1, Diego E. Shalom2 and Marcos A. Trevisan1,2 1 Physics Institute of Buenos Aires (IFIBA) CONICET, Buenos Aires, Argentina, 2 Department of Physics, Universidad de Buenos Aires, Buenos Aires 1428EGA, Argentina. Corresponding author: [email protected] (MAT) March 2, 2018 Speech requires programming the sequence of vocal gestures that produce the sounds of words. Here we explored the timing of this program by asking our participants to pronounce, as quickly as possible, a sequence of consonant-consonant-vowel (CCV) structures appearing on screen. We measured the delay between visual presentation and voice onset. In the case of plosive consonants, produced by sharp and well defined movements of the vocal tract, we found that delays are positively correlated with the duration of the transition between consonants. We then used a battery of statistical tests and mathematical vocal models to show that delays reflect the motor planning of CCVs and transitions are proxy indicators of the vocal effort needed to produce them. These results support that the effort required to produce the sequence of movements of a vocal gesture modulates the onset of the motor plan. I. INTRODUCTION Voluntary movements need preparation before execution [1]. This preparatory phase received extensive attention from the scientific community, whose investigations across species [2,3] helped unveiling the instructions to the effectors before tasks as locomotion [4], arm reaching [5] and vocal production [6,7]. In humans, the readiness potential (RP) is considered a universal signature of planning of volitional acts that has been largely investigated with electroencephalographic techniques [8–10]. Speech requires programming the instructions to drive two acoustically uncoupled effectors: the vocal folds and vocal tract [11]. The folds are a pair of opposed membranes located at the larynx, which leads to a series of configurable cavities ending at the mouth, known as vocal tract. When the larynx is adducted and lung pressure reaches a threshold, the folds oscillate colliding with each other [12]. The resulting periodic interruptions of airflow produce a pressure wave that propagates along the vocal tract. When the tract is open, the sound radiated from the mouth corresponds to a vowel [13]. Consonants are produced in different manners: fricatives are vortex sounds produced by the turbulent passage of air through a constriction in the vocal tract [14]. When lips are narrowed, an [f] is produced; when the tip of the tongue approaches the upper teeth, we hear [s]. Plosives are generated by completely occluding the vocal tract at a given point [15]. When lips are released, we hear [b]; releasing the tip or the body of the tongue produces a [d] or a [g], respectively. From a mechanical point of view, these families of sounds represent vocal gestures of different effort: vowels are produced by smooth and opened tract configurations, fricatives by narrow constrictions and plosives by sharp closures. Recent advances give a cause for great optimism in describing vocal motor control in the language of dynamical systems [5]. As a matter of fact, the avian vocal system presents strong analogies with the human case [16], and birdsong has been fully described in dynamical terms [7,17], leading to devices that generate song controlled by a few physiological variables [18]. What are the dynamical variables driving the human vocal system? The production of speech requires coordination of the larynx, lips, tongue and jaw to produce a gesture, and then control of 'the strength with which the associated gesture (e.g., lip closure) "attempts" to shape vocal tract movements at a given point in time' [19]. These two variables, the gesture and its strength, have been interpreted in dynamical terms [19]. In this theoretical description, the vocal program is executed by coupled neural oscillators; their relative phases represent the sequence of movements of the articulators that define a gesture, and their amplitude represents its strength. Here we hypothesized that the vocal program is sensitive to the strength of the vocal gestures, and that this dependence is revealed in the timing between the instruction and the vocal onset. To explore this hypothesis, we created and analyzed a database of speech structures produced under a simple reaction-time paradigm. II. MATERIALS AND METHODS A. Participants Forty five native Spanish speakers (23 females, age range 20-28, mean age 22), undergraduate and graduate students at the University of Buenos Aires, with normal or corrected-to-normal vision and no speech impairments, completed the experiment. All the participants signed a written consent form. All the experiments described in this paper were approved by the ethics committee Comité de Ética del Centro de Educación Médica e Investigaciones Clínicas 'Norberto Quirno' (CEMIC) qualified by the Department of Health and Human Services (HHS, USA): IRb00001745-IORG 0001315. B. Stimuli and tasks Participants sat in a silent room, 0.7 m away from a 19' monitor. They were asked to pronounce a sequence of consonant-consonant-vowel (CCV) structures, as soon as they appeared at the center of the screen. The CCVs were recorded with a commercial microphone at approximately 0.2 m of the mouth. After pronouncing a CCV, the participant was instructed to press a key. A blank screen preceded the following CCV, which appeared after a random time between 0.5 and 2 s. The procedure was repeated until completing 3 rounds of the complete CCV set. All the presentations were randomized. The experiment was coded in Python, using Pyaudio and PsychoPy [20] libraries to ensure robust audio-video synchronization (30 ms maximum offset between visual presentation and audio recording). CCVs were formed by combinations of fricatives and plosives, followed by the vowel [a]. We used the most common Spanish fricatives [21,22] [f, s, x], pronounced as in face, stand and loch, and the most common Spanish plosives [b, p, d, t, g, k] [21], pronounced as in bay, pay, die, tie, gray and cray. Plosives come in voiced-unvoiced pairs [b, p], [d, t] and [g, k]. The tract anatomy is identical for each pair, but folds are active only for the first components. Since plosives are recognized through the sound produced at the occlusion release, unvoiced ones leave no acoustical traces during occlusion. On the contrary, the folds' fundamental frequency serves to spectrally mark vocal onset during occlusion in voiced plosives. For this reason, CCVs starting with plosives were reduced to the voiced ones [b, d, g]. We presented the 6 possible combinations of fricatives (fja, fsa, jfa, jsa, sfa, sja), the 9 possible plosive-fricative combinations (bfa, bja, bsa, dfa, dja, dsa, gfa, gja, gsa), a selection of 11 fricative- plosive combinations (fga, fba, fpa, jba, jca, jga, jda, sba, sca, sga, sta) and 11 combinations of plosives (bca, bda, bga, bta, dba, dca, dga, dta, gba, gda, gta). This set of 37 CCVs allowed keeping a short (mean 20 minutes), single-session experiment per participant. The number of recorded samples was 4995 = 45 subjects × 3 trials × (6 fricative-fricative + 9 plosive-fricative +11 fricative-plosive + 11 plosive-plosive) combinations. C. Data selection Stage 1. Participants were instructed to pronounce the CCVs clearly, as soon as possible after visual presentation, and they were not allowed to correct their utterances. These requirements produced a considerable amount of errors during the experiment. We discarded audio files that a. contained more than one attempt to produce the CCV and b. did not match with the presented CCV. This left us with 3296 CCVs. Stage 2. For each audio file that passed stage 1, we measured three timing quantities: the delay Δ between visual presentation and voice onset, the transition  between consonants and duration T of the CCV. Audio files were inspected with Praat [23]. Delays Δ were measured directly from the spectrogram. Transitions are defined as the time  during which the vocal tract changes its configuration from the first to the second consonant. Since vocal gestures producing the same speech content are not necessarily unique [24], we extracted by inspection the most robust features of each CCV across trials and participants, and therefore only selected the samples that strictly met the following spectral signatures for consonantal transitions: a. Fricative-fricative combinations. As noisy turbulent sounds, the spectral signature of a fricative is a dark spot, such as the ones shown in the left panel of Fig. 1A for [f] and [s]. Transitions are defined by the continuous passage from the first to the second constriction, therefore presenting a mixture of spectral components from both consonants. b. Plosive-fricative combinations. Transitions are characterized by the interval that goes from releasing the occlusion of the plosive to the formation of a constriction to generate the fricative. Spectrally, the transition is described by the voiced sound structure produced after the plosive, and before the formation of the purely noisy fricative spot (Fig. 1B, left). c. Fricative-plosive combinations. During these transitions, the vocal tract evolves from a constriction at one point to an occlusion at another one. Transitions are spectrally defined as the interval between the abrupt end of the fricative and the release of the plosive into the vowel [a], as shown in the left panel of Fig. 1C. d. Plosive-plosive combinations. Vocal tract passes from an occlusion at a given point of the tract to another one at a different location. Transitions, as shown in the representative case of the left panel of Fig. 1D, go from the release of the first plosive characterized by a voiced sound structure, to the release of the second one within the vowel [a]. A pool of 1958 CCVs was successfully classified into these spectral categories. Four subjects were discarded for presenting systematical spectral differences from the described categories. Stage 3. We finally discarded data that passed stage 2 for which delays Δ and transitions '=/T were greater than two standard deviations from the speaker mean values. A final timing dataset from 1519 CCVs from N=41 participants was used to perform the analyses presented in this work. D. CCV frequency None of the CCVs are Spanish words. We counted the intra-word appearances of each CCV in a large Spanish corpus [25], and assigned the appearances per million words as the frequency of each CCV. Since the range of frequency values was large (from 0 to roughly 105 apmw), in Fig. 2 we show a discretization of this range in low, medium or high frequency values. E. Vocal model The main ingredients of the folds' dynamics are caught by the following equation of motion, known as the flapping model [26]: m z=−k (z) z−b( z, z) z+ ps δ+2Γ z z0+δ /2+z+Γ z (1) In Fig. 3A we sketch its elements. The variable z represents the transverse displacement of the folds from the pre-phonatory position z0. The tissue is described by nonlinear elastic k and dissipative b functions. The last term represents the energy transferred to the folds by lung pressure ps, which depends on the parameters δ and Γ that define the folds' configuration. Dynamical analysis of Eq. 1 revealed that pressure ps beyond a threshold creates oscillations through a Hopf bifurcation [12] and that the threshold increases with glottal abduction z0 [27]. Oscillations make the folds collide, interrupting the airflow at the vocal tract entrance and producing pressure waves p that propagate along a vocal tract. The tract cross-section A(x,t) is shown in Fig. 3B, where x represents the distance from the glottis (x=0) to the mouth (x=L). During a general sequence of vowels and plosive consonants, A(x,t) was described by Story [28]: A (x ,t)=π/4 [Ω(x)+q1(t)ϕ1(x)+q2(t)ϕ2( x)]2∏ n k=1 [1−ck(x)m(t−t k)] (2) The first factor represents the vowel substrate, with empirical functions Ω, φ1 and φ2 obtained from an orthogonal decomposition calculated from MRI data [29]. Any sequence of vowels is generated by the evolution of q1 and q2. Plosive consonants are generated through the bell-shaped functions ck and m. These functions reach the value 1 around x=xk and t=0 respectively, producing a specific occlusion A(xk,tk)=0. The spatial functions ck represent the anatomy of the occlusion for each consonant, and the time function m represents its activation-deactivation. F. CCV synthesis Eqs. 1 and 2 were solved with MATLAB. Numerical integration of Eq. 2 was performed through a standard Runge-Kutta algorithm at a sampling rate of 44.1 kHz. A wave-reflection model [30,31] was used to simulate propagation of sound along the vocal tract, which is approximated by a series of N=44 tubes of cross-section Ai, 1≤i≤N. Sound propagation is solved by splitting an incoming sound wave pi into a reflected and a transmitted wave at each interface, with reflection and transmission coefficients ri,j=(Ai-Aj)/(Ai+Aj) and ti,j=1-ri,j (j=i±1) [32]. The sound radiated from the mouth is proportional to the pressure at the last tube, p44. The resulting time series was converted to wav audio format. To synthesize CCVs for voiced plosives, we proceeded as follows: 1. Subglottal pressure ps was set at a normal speech value such that Eq. 1 generated oscillations of the folds with a pitch around 100 Hz, from t=0 to voice offset at t=Tf. The rest of the parameters can be found elsewhere [11]. 2. Linear functions q1(t)=4t/T and q2(t)=t/T [26] were used in Eq. 2 to drive the tract from a neutral tract (q1,q2)=(0,0) at t=0 to the shape of the vowel [a] (q1,q2)=(4,1) at t=T, and maintained there until t=Tf. 3. While the tract evolves towards the vowel [a], two occlusions take place at t1 and t2, activated through Gaussian functions m of width 0.14 s [28], separated by τ=t2-t1. We used the mean value of τ across CCVs and participants. The spatial functions ck reach the value 1 at xkL for [b], xk9/10L for [d] and xk7/10L for [g]. Functional forms and parameter values for the functions ck and m are reported elsewhere [28,33]. III. RESULTS We investigated the vocal planning prior to the utterance of short speech structures. For that sake, we recorded 45 speakers while pronouncing a set of consonant-consonant-vowel structures (CCV) as soon as they appeared on a computer screen. We used combinations of fricatives [f, s, x] and plosives [b, p, d, t, g, k] followed by the vowel [a]. Examples of these structures are fsa, gfa, fga and dca, which are also representatives of fricative-fricative, plosive-fricative, fricative-plosive and plosive-plosive combinations respectively. For each audio file, we measured three timing variables: the delay between visual presentation and phonation onset (Δ), the duration of the CCV (T) and the duration of the transition between the consonants (τ). In the left panels of Fig. 1 we show these timing variables for representative cases of the fricative-fricative (A), fricative-plosive (B), plosive-fricative (C) and plosive-plosive (D) types. After selecting the audio files that met the spectral criteria for each family type (see section II C), we collected timing data from 1519 CCVs produced by 41 subjects. How are these timing variables related? To compare dynamics across participants with different speech rates, we normalized the transitions to the duration T of the CCV, τ'=τ/T. In the right panels of Fig. 1 we plot Δ vs. τ' for the different families of consonantal combinations. No correlations were found between delays and transitions for fricatives (Fig. 1A, r=-0.27, p=0.6) or combinations of fricatives and plosives (Fig. 1B, r=-0.29 p=0.38 and Fig. 1C, r=0.15, p=0.7). Interestingly, however, a positive correlation emerged for combinations of plosive consonants (Fig. 1D, r=0.68, p<0.02). We next explored how these timing variables relate to relevant motor and planning variables. FIG. 1. Timing variables for the different consonantal transitions. Left panels show onset delays Δ and consonantal transitions τ measured from sound and spectrogram, for representative cases of the fricative-fricative (A), fricative-plosive (B) plosive- fricative (C) and plosive-plosive (D) types. In the right panels we show that vocal onset delays Δ are not significantly correlated with the normalized transitions τ'=τ/T for the 6 fricative-fricative (A), 11 fricative-plosive (B) and 9 plosive-fricative (C) combinations. On the contrary, timing variables are positively correlated for the 11 plosive-plosive combinations (D). A. Delays and motor planning Vocal onset delays represent the integration of reading, memory retrieval and vocal planning tasks, followed by articulatory movements produced before phonation. We first analyzed the effects of each of these contributions. Reading. Reading task involve visualization and recognition of three common Spanish phonemes of identical visual size. Given the short length of the strings, we assumed that this effect is constant across CCVs. Memory retrieval. We explored two memory variables. Long term effects were investigated by computing the frequency of occurrence of the CCVs in Spanish, shown in Fig. 2 (see details in section IV D). Since our participants completed three rounds of the experiment, trial number was used to investigate short term memory effects. We submitted timing data to ANOVA with frequency and trial as independent variables (Table I). The analysis revealed effects of frequency on Δ and τ' for all the consonant combinations, except for the case of plosive-plosive type. This merely reflects that these are very unusual combinations in Spanish, and therefore no long term memory effects were detected. Trial number, on the contrary, has a significant effect on Δ and no effect on τ', for any consonantal combination. A follow up of this analysis revealed that this effect on Δ was accounted for by an increase in Δ for trial 1 (mean 0.94±0.01 s) than for trials 2 and 3 (mean 0.85±0.01 s), (t=8.13, df=1176, p<10-14). We therefore considered trial 1 as a training set and used timing data from trials 2 and 3 for the following analyses. Articulatory kinematics. Since Δ was measured from visual presentation to vocal onset, articulatory kinematics that normally precede phonation [34,35] are included within this time period (e.g. lip closure before folds' activation for an initial [b]). To study the effect of movements prior to sound onset, submitted our data to ANOVA using the first consonant of the CCV as independent variable, restricting the study to initial plosives [b], [d] and [g]. The production of these consonants before phonation require articulation of the lips, tongue tip and tongue body movements respectively. This analysis revealed no significant effects of initial plosives on Δ (p=0.323, F=1.13, df=2), which allowed us to assume that the period of articulation previous to phonation is constant across CCVs. Due to the nature of the experiments performed here, vocal planning cannot be explored directly. Instead, we analyzed the different tasks performed by the participants from CCV visualization to phonation. We found that, for plosive-plosive combinations, the tasks of reading, memory and prephonatory movements can be assumed constant across CCVs. We therefore assumed that variations of Δ across CCVs account for differences in the vocal planning. FIG. 2. Frequency of occurrence of consonants in CCVs. Rows and columns correspond to the first and the second CCV consonants respectively. Colors represent low (blue), medium (light blue) and high (yellow) frequency of occurence values of the these combinations in Spanish. Occurrences of fricative-fricative and fricative-plosive combinations present high variability, while only one plosive-plosive combination presents a high occurrence level. B. Consonantal transitions and vocal effort Under the simplistic hypothesis that effortless gestures are produced faster than harder ones, we could roughly assume that transitions τ' reflect the effort needed to produce a CCV. We first note that, consistently with this hypothesis, the right panels Fig. 1 show that the shorter transitions correspond to fricatives (τ'=0.124±0.002), which require relatively low efforts to produce constrictions in the tract; longer transitions were found in the production of plosive- fricative (τ'=0.163±0.003) or fricative-plosive combinations ('=0.280±0.003), and these are in turn shorter than transitions between plosives ('=0.389±0.003), which require the largest efforts to completely occlude the vocal tract twice. To further formalize this notion of effort, we propose a physically-based model that represents the effort E exerted by the tract and the folds during speech. The expression reads: E=E tract+ Efolds=∫ ∫ A (x ,t )− AΩ(x)dxdt +E 0 (3) T L 0 0 The shape of the vocal tract is mathematically described by its cross-section A(x,t), with x the distance from the glottis (x=0) to the mouth (x=L), as shown in the sketch of Fig. 3B. Building on a previous mass-spring model [27], the first term of Eq. 3 represents the elastic force exerted by the departures of the tract from a relaxed configuration AΩ(x) along an utterance of duration T. During a CCV of plosive consonants, two occlusions separated by  occur while the tract is evolving towards the vowel. This is described in Eq. 2 (section II E), where plosives are spatially described by functions ck(x) that model the extension and location of the occlusion, and temporally described by functions m(t) that control the activations of closures. Using the parameters reported for consonantal activations [19,28], efforts computed using each CCV transition (in a range that goes from DGA=0.33±0.03 s to GTA=0.45±0.04 s, shown in Fig. 3C) display the same organization as the efforts computed for all the CCVs using the population mean (=0.40±0.04 s) for all CCVs. To build confidence on the pertinence of these temporal parameters, we used Eqs. 1 and 2 to generate synthetic CCVs with the activations reported (see section II F) and fixed  mean value, obtaining audio files with the same spectral features and speech content than the experimental records (Supplementary Materials). Tract effort does not explicitly depend on , but instead on the specific consonants forming the CCV and their order of appearance. The dependence on the particular CCV consonants stem from the anatomical differences between them. For instance, the function ck(x) that models the extension of occlusion is wider for [d] than for [b], accounting for the greater extent where the constriction area is zero when it is performed with the tongue than with the lips; a greater effort is then needed to produce a [d] than a [b]. The dependence of the effort with the order of appearance of the consonants is due to the fact that consonants occur while the vocal tract is evolving towards the vowel [a]. For instance, bda requires less effort than dba. This asymmetry is explained because in dba, the lips closure for [b] occurs closer to the [a], for which the mouth is fully opened, requiring a larger vocal tract deformation that corresponds to a larger effort. Transitions ' systematically reflect this asymmetry: 'DGA<'GDA (t=- 2.9, df=49.8, p<0.01); 'BGA<'GBA (t=-2.13, df=86.9, p<0.05) and 'BDA<'DBA (t=-0.71, df=136.5, p=0.48). The second term in Eq. 3 refers to laryngeal efforts. Vibration of the vocal folds has an on/off function in speech. Within our set of CCVs, vibration is either active during the whole vocalization (for the subset bda, dba, bga, gba, dga, gda) or inactivated during the second (unvoiced) plosive (subset bca, bta, dca, gta, dta). This second group involves a devoicing effort by glottal abduction. We model Efolds in Eq. 3 as a constant that takes the value E0=0 for the first group and E0>0 for the second one. The distribution of CCVs along E varies with E0. For instance, a sufficiently large value of E0 separates one group from the other, eventually producing a bimodal distribution. We set the value of E0 that matched the experimental distribution of CCVs along ' (vertical axis of Fig. 3C) to the theoretical distribution along E (horizontal axis of Fig. 3C). We obtained values of E0 between 7% and 11% of the mean vocal tract effort (for distributions between 4-8 bins). The effort model of Eq. 3 represents a compromise between mathematical simplicity and vocal biomechanics, accounting for both tract reconfigurations and control of folds, which are the main gestures necessary to generate speech. Although effort does not depend on ' within the range of measured values, Fig. 3C shows that both variables are significantly correlated (R=0.84, p<0.001), reflecting that consonantal transitions are indeed shorter when the global gesture requires less effort. This implies that ' is a good proxy for the effort needed to produce the CCVs. FIG. 3. Elements of the articulatory model used to calculate vocal effort. A. Vocal tract movements, described by its cross-section A(x,t), require an effort proportional to its deformations with respect to a relaxed configuration. B. Elements of the vocal folds model of Eq. 1. Abduction effort is needed to separate the folds, increasing z0 in Eq. 1 to deactivate oscillations during unvoiced plosives. This is modeled as a constant E0 for the group (bca, bta, dca, gta, dta) shown in green. C. The transitions  measured on the recorded CCVs are positively correlated with vocal effort E computed with Eq. 3. The distributions of CCVs with E and  are shown along the horizontal and vertical axis respectively, for E0=9% of the mean tract effort. Taken together, these results show that the timing variables measured directly from speech recordings are related to planning and motor variables. For combinations of plosive consonants, we found that differences in Δ account for differences in the motor planning and that ' are proportional to the effort needed to produce the global CCV gesture. IV. CONCLUSIONS AND DISCUSSION Here we hypothesized that the vocal program is sensitive to the strength of the vocal gestures, and that this dependence is revealed in the time required to activate vocal production. To explore this, we created and analyzed a database of speech structures produced under a simple reaction-time paradigm. Our participants were instructed to pronounce a set of consonant-consonant-vowel structures (CCV) as soon as they appeared on screen. We measured two timing quantities from the recorded audio files: the delay between visual presentation and vocal onset, and the duration of the transition between consonants, normalized to the duration of the CCV. Using a battery of tools that includes statistical tests and nonlinear modeling of the vocal system, we found that 1. these variables are positively correlated for CCVs formed with plosive consonants. Then, we show the plausibility that 1. variations in delays account for variations in vocal planning and 2. consonantal transitions reflect the vocal effort required to produce the CCVs. Overall, these results support the idea that the time required to plan short speech structures is proportional to the effort needed to produce them. The fact that no correlations were found for the rest of the consonantal combinations can be due to multiple factors. In particular, the memory effects detected on these combinations suggests a probable interaction between motor planning and memory tasks. Our findings are indeed compatible with electroencephalographic measurements of the readiness potential (RP), defined as the buildup of electrical potential beginning about one second before cued or self-paced movement onset. This universal signature of planning and initiation of volitional acts [8] has been largely explored, and previous works reported that the level of force and speed of a motor task can be detected from these cortical potentials [10]. Moreover, differences in RP onset time account not only for initial movements but also for a sequence of movements [36]. Evidence is accumulating towards the vision of motor cortex as a dynamical system that generates and control movements [5]. The particular case of the vocal motor program has been extensively studied and successfully described in dynamical terms for songbirds [7]. Although strong analogies connect the avian and human vocal systems, descriptions of the human motor program in terms of dynamical systems received considerably less attention. One exception is the approach proposed by Goldstein and collaborators [19], whose theoretical model represents the vocal motor program as a set of coupled neural oscillators controlling the vocal articulators. The relative phases of the oscillators and their amplitudes configure a vocal gesture and its strength. The model of vocal effort presented here presents a plausible quantification of this notion of vocal strength, supporting the hypothesis that the timing of the motor program is sensitive to the strength of the vocal gestures. It is well established that brain motor areas activate not only for voice production, but also to assist other non-motor tasks as speech perception [37]. More interestingly perhaps, these motor maps are also used in the case of purely mental operations such the inner speech, i.e. the voice that "narrates the words we read or the conversations we imagine" [38]. In fact, as happens with speech itself, inner speech is associated with an efference copy with detailed auditory properties [38]. In light of these results, we plan to capitalize on the simple experimental paradigm presented here to explore vocal motor signatures during silent lecture. Acknowledgements The research reported in this work was partially funded by the Concejo Nacional de Investigaciones Científicas y Técnicas, CONICET; The University of Buenos Aires UBA and NIH through R01- DC-012859. References [1] M. S. Graziano, Neuron 71, 387 (2011). [2] W. A. Friedman, L. M. Jones, N. P. Cramer, E. E. Kwegyir-Afful, H. P. Zeigler, and A. Keller, J [3] [4] Neurophysiol 95, 1274 (2006). A. Riehle and J. Requin, J. Neurophysiol. 6, (1989). G. S. Stent, W. B. Kristan, W. O. Friesen, C. a Ort, M. Poon, and R. L. Calabrese, Science 200, 1348 (1978). [5] M. M. Churchland, J. P. Cunningham, M. T. Kaufman, J. D. Foster, P. Nuyujukian, S. I. Ryu, and K. V Shenoy, Nature 487, 51 (2012). [6] R. G. Alonso, M. a. Trevisan, A. Amador, F. Goller, and G. B. Mindlin, Front. Comput. Neurosci. 9, 1 (2015). [7] A. Amador, Y. S. Perl, G. B. Mindlin, and D. Margoliash, Nature 495, 59 (2013). [8] [9] A. Schurger, J. D. Sitt, and S. Dehaene, Proc. Natl. Acad. Sci. 109, E2904 (2012). V. Siemionow, G. H. Yue, V. K. Ranganathan, J. Z. Liu, and V. Sahgal, Exp. Brain Res. 133, 303 (2000). [10] M. Jochumsen, I. K. Niazi, N. Mrachacz-Kersting, D. Farina, and K. Dremstrup, J. Neural Eng. 10, (2013). [11] R. Laje, T. Gardner, and G. B. Mindlin, Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys. 64, 056201/1 (2001). [12] [13] J. C. Lucero and L. L. Koenig, J. Acoust. Soc. Am. 117, 1362 (2005). I. R. Titze, J. Acoust. Soc. Am. 83, 1536 (1988). [14] M. H. Krane, J. Acoust. Soc. Am. 118, 410 (2005). [15] B. H. Story and K. Bunton, J. Speech. Lang. Hear. Res. 53, 1514 (2010). [16] H. P. Zeigler and P. Marler, Neuroscience of Birdsong (Cambridge Univ Pr, 2008). [17] M. Trevisan, G. Mindlin, and F. Goller, Phys. Rev. Lett. 96, 1 (2006). [18] [19] [20] [21] E. M. Arneodo, Y. S. Perl, F. Goller, and G. B. Mindlin, PLoS Comput. Biol. 8, e1002546 (2012). L. Goldstein, D. Byrd, and E. Saltzman, in Action to Laguage via Mirror Neuron Syst. (Cambridge University Press, 2006), pp. 215–249. J. W. Peirce, J. Neurosci. Methods 162, 8 (2007). J. I. Hualde, A. Olarrea, and E. O'Rourke, The Handbook of Hispanic Linguistics (2012). [22] A. M. Borzone de Manrique and M. I. Massone, J. Acoust. Soc. Am. 69, 1145 (1981). [23] P. Boersma and D. Weenink, (2013). [24] M. F. Assaneo, M. A. Trevisan, and G. B. Mindlin, PLoS One 8, e80373 (2013). [25] F. Cuetos, M. Glez-Nosti, A. Barbón, and M. Brysbaert, Psicológica 32, 133 (2011). [26] M. F. Assaneo, J. I. Nichols, and M. A. Trevisan, PLoS One 6, e28317 (2011). [27] R. Kirchner, An Effort Based Approach to Consonant Lenition (Routledge, 2013). [28] B. H. Story, J. Acoust. Soc. Am. 117, 3231 (2005). [29] B. H. Story and I. R. Titze, J. Phon. 26, 223 (1998). [30] M. M. Sondhi, J. Acoust. Soc. Am. 55, 1070 (1974). [31] J. Flanagan, J. Allen, and M. Hasegawa-Johnson, Speech Analysis, Synthesis and Perception (2008). [32] B. H. Story, Physiologically-Based Speech Simulation Using an Enhanced Wave-Reflection Model of the Vocal Tract, University of Iowa, 1995. [33] M. F. Assaneo, J. Sitt, G. Varoquaux, M. Sigman, L. Cohen, and M. A. Trevisan, Neuroimage 141, 31 (2016). [34] K. E. Bouchard, N. Mesgarani, K. Johnson, and E. F. Chang, Nature 495, 327 (2013). [35] D. F. Conant, K. E. Bouchard, M. K. Leonard, and E. F. Chang, J. Neurosci. 2382 (2018). [36] M. Simonetta, M. Clanet, and O. Rascol, Electroencephalogr. Clin. Neurophysiol. Evoked Potentials 81, 129 (1991). [37] [38] J. M. Correia, B. M. B. Jansma, and M. Bonte, J. Neurosci. 35, 15015 (2015). T. J. Whitford, B. N. Jack, D. Pearson, O. Griffiths, D. Luque, A. W. F. Harris, K. M. Spencer, and M. E. Le Pelley, Elife 1 (2017). Tables TABLE I. ANOVA analysis with frequency of appearance of the CCVs in Spanish and trial number as independent factors. We also performed the tests using the absolute frequency values, with equivalent results. Frequency (low, medium, high) Trial number p<10-5, F=9.54, df=2 p=0.004, F=8.37, df=2 p<10-5, F=42.6, df=2 p=0.4, F=0.89, df=2 fricative-fricative fricative-plosive p=6×10-4, F=7.51, df=2 p=0.4, F=0.93, df=2 plosive-fricative p=0.002, F=9.93, df=1 p=0.12, F=2.1, df=2 Δ ' plosive-plosive p=0.62, F=0.24, df=1 p=0.59, F=0.54, df=2 Supplementary Materials supplementary.zip contains the pool recorded of CCVs in wav format from our experiment, marked with the timing variables (in Praat format).
1607.01706
3
1607
2017-01-10T04:18:23
Optimal Trajectories of Brain State Transitions
[ "q-bio.NC" ]
The complexity of neural dynamics stems in part from the complexity of the underlying anatomy. Yet how the organization of white matter architecture constrains how the brain transitions from one cognitive state to another remains unknown. Here we address this question from a computational perspective by defining a brain state as a pattern of activity across brain regions. Drawing on recent advances in network control theory, we model the underlying mechanisms of brain state transitions as elicited by the collective control of region sets. Specifically, we examine how the brain moves from a specified initial state (characterized by high activity in the default mode) to a specified target state (characterized by high activity in primary sensorimotor cortex) in finite time. Across all state transitions, we observe that the supramarginal gyrus and the inferior parietal lobule consistently acted as efficient, low energy control hubs, consistent with their strong anatomical connections to key input areas of sensorimotor cortex. Importantly, both these and other regions in the fronto-parietal, cingulo-opercular, and attention systems are poised to affect a broad array of state transitions that cannot easily be classified by traditional notions of control common in the engineering literature. This theoretical versatility comes with a vulnerability to injury. In patients with mild traumatic brain injury, we observe a loss of specificity in putative control processes, suggesting greater susceptibility to damage-induced noise in neurophysiological activity. These results offer fundamentally new insights into the mechanisms driving brain state transitions in healthy cognition and their alteration following injury.
q-bio.NC
q-bio
Optimal Trajectories of Brain State Transitions Shi Gua,b, Richard F. Betzelb, Marcelo G. Mattarc, Matthew Cieslakd, Philip R. Deliod,e, Scott T. Graftond, Fabio Pasqualettif, Danielle S. Bassettb,g,∗ aApplied Mathematics and Computational Science, University of Pennsylvania, Philadelphia, PA, 19104 USA bDepartment of Bioengineering, University of Pennsylvania, Philadelphia, PA, 19104 USA cPrinceton Neuroscience Institute, Princeton University, Princeton NJ 08544 dDepartment of Psychological and Brain Sciences, University of California, Santa Barbara, CA, 93106 USA eNeurology Associates of Santa Barbara, Santa Barbara, CA, 93105 USA fDepartment of Mechanical Engineering, University of California, Riverside, CA, 92521 USA g Department of Electrical & Systems Engineering, University of Pennsylvania, Philadelphia, PA, 19104 USA 7 1 0 2 The complexity of neural dynamics stems in part from the complexity of the underlying anatomy. Yet how white matter structure constrains how the brain transitions from one cognitive state to another remains unknown. Here we address this question by drawing on recent advances in network control theory to model the underlying mechanisms of brain state transitions as elicited by the collective control of region sets. We find that previously identified attention and executive control systems are poised to affect a broad array of state transitions that cannot easily be classified by traditional engineering-based notions of control. This theoretical versatility comes with a vulnerability to injury. In patients with mild traumatic brain injury, we observe a loss of specificity in putative control processes, suggesting greater susceptibility to neurophysiological noise. These results offer fundamental insights into the mechanisms driving brain state transitions in healthy cognition and their alteration following injury. Keywords: network neuroscience, control theory, traumatic brain injury, cognitive control, diffusion imaging Abstract n a J 0 1 ] . C N o i b - q [ 3 v 6 0 7 1 0 . 7 0 6 1 : v i X r a Introduction The human brain is a complex dynamical system that transitions smoothly and continuously through states that directly support cognitive function (Deco et al., 2011). Intuitively, these trajectories can map out the mental states that our brain may pass through as we go about the activities of daily living. In a mathematical sense, these transitions can be thought of as trajectories through an underlying state space (Shenoy et al., 2011; Freeman, 1994; Gu et al., 2016). While an understanding of these trajectories is critical for our understanding of cognition and its alteration following brain injury, fundamental and therefore generalizable mechanisms explaining how the brain moves through states have remained elusive. One key challenge hampering progress is the complexity of these trajectories, which stems in part from the architectural complexity of the underlying anatomy (Hermundstad et al., 2011, 2013a, 2014a). Different components (neurons, cortical columns, brain areas) are linked with one another in complex spatial patterns that enable diverse neural functions (Rajan et al., 2016; Fiete et al., 2010; Levy et al., 2001). These structural interactions can be represented as a graph or network, where component parts form the nodes of the network, and where anatomical links form the edges between nodes (Bullmore & Sporns, 2009). The architecture of these networks displays heterogenous features that play a role in neural function (Medaglia et al., 2015), development (Di Martino et al., 2014), disease (Braun et al., 2015), and sensitivity to rehabilitation (Weiss et al., 2011). Despite these recent discoveries, how architectural features constrain neural dynamics in any of these phenomena is far from understood. One simple and intuitive way to formulate questions about how neural dynamics are constrained by brain network architecture is to define a state of the brain by the 1 × N vector representing magnitudes of neural activity across N brain regions, and to further define brain network architecture by the N × N adjacency matrix representing the number of white matter streamlines linking brain regions (Gu et al., 2015). Building on these two definitions, we can ask how the organization of the white matter architecture constrains the possible states in which the brain can or does exist (Durstewitz & Deco, 2008; Hansen et al., 2015). Moreover, building on decades of cognitive neuroscience research that have carefully delineated the ∗Corresponding author Email address: [email protected] (Danielle S. Bassett) Preprint submitted to NeuroImage October 26, 2018 Figure 1: Conceptual Schematic. (A) Diffusion imaging data can be used to estimate connectivity from one voxel to any other voxel via diffusion tractography algorithms. (B) From the tractography, we construct a weighted network in which N = 234 brain regions are connected by the quantitative anisotropy along the tracts linking them (see Methods). (C) We study the optimal control problem in which the brain starts from an initial state (red) at time t = 0 and uses multi-point control (control of multiple regions; blue) to arrive at a target state (yellow) at time t = T . role of regional activation in cognitive functions (Gazzaniga, 2013; Szameitat et al., 2011; Alavash et al., 2015), we can then map brain states to cognitive processes, and extend our question to: how does the organization of white matter architecture constrain cognitive states (Hermundstad et al., 2013a, 2014b), and the processes that enable us to move between those cognitive states (Cocchi et al., 2013)? To address these questions, we draw on recent advances in network control theory (Pasqualetti et al., 2014) to develop a biologically-informed mathematical model of brain dynamics from which we can infer how the topology of white matter architecture constrains how the brain may affect (or control ) transitions between brain states. Within this model, we examine finite-time transitions (from initial to target state) that are elicited via the collective control of many regions, consistent with the collective dynamics observed to support cognition (Salvador et al., 2005; Meunier et al., 2009; Power et al., 2011; Yeo et al., 2011) and action (Bassett et al., 2011b, 2013, 2015). A natural choice for an initial state is the brain's well-known baseline condition, a state characterized by high activity in the precuneus, posterior cingulate, medial and lateral temporal, and superior frontal cortex (Raichle, 2015; Raichle & Snyder, 2007; Raichle et al., 2001). While potential transitions from this default mode are myriad, we focus this first study on examining transitions into target states of high activity in sensorimotor cortex: specifically the extended visual, auditory, and motor cortices. These states represent the simplest and most fundamental targets to transition from the default mode: for example, transitioning from the default mode to visual states might represent an immediate response to a surprising stimuli. Similarly, the transition from the default mode to motor states might represent the simple transition from rest to action. Moreover, these transitions are of particular interest in many clinical disorders including stroke (Carter et al., 2012) and traumatic brain injury (Nudo, 2006; Lee et al., 2011) where the cognitive functions performed by these target areas are often altered, significantly effecting quality of life (Kalpinski et al., 2013). Using network control theory, we examine the optimal trajectories from an initial state (composed of high activity in the default mode system) to target states (composed of high activity in sensorimotor systems) with finite time and limited energy. In this optimal control context, we investigate the role of white matter connectivity between brain regions in constraining 2 00.050.100.150.200.250.300.35RegionRegion0TABCInital StateTarget StateControl RegionsInputConnectivity Strength12341001234100 dynamic state transitions by asking three interrelated questions. First, we ask which brain regions are theoretically predicted to be most energetically efficient in eliciting state transitions. Second, we ask whether these state transitions are best elicited by one of three well-known control strategies commonly utilized in mechanical systems (Gu et al., 2015). Third, we ask how specific each region's role is in these state transitions, and we compare this specificity between a group of healthy adults and a group of patients with mild traumatic brain injury. In particular, the inclusion of this clinical cohort enables us to determine whether widespread injury leads to a decrement in the healthy network control profiles, thus requiring greater energy for the same functions, or an enhancement of the healthy network control profiles at the cost of a more fragile system, overly sensitive to external perturbations. Together, these studies offer initial insights into how structural network characteristics constrain transitions between brain states, and predict their alteration following brain injury. To address these questions, we build structural brain networks from diffusion spectrum imaging (DSI) data acquired from 48 healthy adults and 11 individuals with mild traumatic brain injury (Fig. 1A). We perform diffusion tractography on these images to estimate the quantitative anisotropy along the streamlines linking N = 234 large-scale cortical and subcortical regions extracted from the Lausanne atlas (Cammoun et al., 2012; Daducci et al., 2012). We summarize these estimates in a weighted adjacency matrix whose entries reflect the number of streamlines connecting different regions (Fig. 1B). We then define a model of brain state dynamics informed by the weighted adjacency matrix, and we use this model to perform a systematic study of the controllability of the system. This construction enables us to examine how structural network differences between brain regions impact their putative roles in controlling transitions between cognitive states (Fig. 1C). Materials and Methods Data Acquisition and Brain Network Construction Diffusion spectrum images (DSI) were acquired from 59 human adults with 72 scans in total, among which 61 scans were acquired from 48 healthy subjects (mean age 22.6± 5.1 years, 24 female, 2 left handed) and 11 were acquired from individuals with mild traumatic brain injury (Cieslak & Grafton, 2014)(mean age 33.8 ± 13.3 years, 4 female, handedness unclear). All participants volunteered with informed written consent in accordance with the Institutional Review Board/Human Subjects Committee, University of California, Santa Barbara. Deterministic fiber tracking using a modified FACT algorithm was performed until 100, 000 streamlines were reconstructed for each individual. Consistent with previous work (Bassett et al., 2010, 2011a; Hermundstad et al., 2013b, 2014a; Klimm et al., 2014; Gu et al., 2015; Muldoon et al., 2016a,b; Sizemore et al., 2015), we defined structural brain networks from the streamlines linking N = 234 large-scale cortical and subcortical regions extracted from the Lausanne atlas (Hagmann et al., 2008). We summarize these estimates in a weighted adjacency matrix A whose entries Aij reflect the structural connectivity (quantitative anisotropy) between region i and region j (Fig. 1A). See SI for further details. Network Control Theory Next, we consider the general question of how the brain moves between different states, where a state is defined as a pattern of activity across brain regions or voxels. In particular, we are interested in studying how the activity in individual brain regions affects the trajectory of the brain as it transitions between states; here, we define a trajectory as a set of states ordered in time. To address this question, we follow (Gu et al., 2015; Muldoon et al., 2016a; Betzel et al., 2016) by adopting notions from the emerging field of network control theory, which offers a theoretical framework for describing the role of network nodes in the control of a dynamical networked system. Network control theory is predicated on the choice of both a structural network representation for the system, and a prescribed model of node dynamics. In the context of the human brain, a natural choice for the structural network representation is the graph on N brain regions whose ijth edge represents the QA between node i and node j. The choice for the model of node dynamics is perhaps less constrained, as many models are available to the investigator. These models range in complexity from simple linear models of neural dynamics with few parameters to nonlinear neural mass models with hundreds of parameters (Gu et al., 2015; Muldoon et al., 2016a). In choosing a model of neural dynamics to employ, we consider multiple factors. First, although the evolution of neural activity acts as a collection of nonlinear dynamic processes, prior studies have demonstrated the possibility of predicting a significant amount of variance in neural dynamics as measured by fMRI through simplified linear models (Gal´an, 2008; Honey et al., 2009; Gu et al., 2015). On the basis of this literature, we employ a simplified noise-free linear continuous-time and time-invariant network model (1) where x : R≥0 → RN describes the state of brain regions over time, and A ∈ RN×N is a symmetric and weighted adjacency matrix. The diagonal elements of the matrix A satisfy Aii = 0. The input matrix BK identifies the control nodes K in the brain, where K = {k1, . . . , km} and x(t) = Ax(t) + Bu(t), (2) and ei denotes the i-th canonical vector of dimension N . The input uK : R≥0 → Rm denotes the control strategy. Intuitively, this model enables us to frame questions related to brain state trajectories in a formal mathematics. Moreover, it allows us BK = [ek1,··· , ek2] 3 to capitalize on recent advances in network control theory (Pasqualetti et al., 2014) to inform our understanding of internal cognitive control (Gu et al., 2015; Betzel et al., 2016) and to inform the development of optimal external neuromodulation using brain stimulation (Muldoon et al., 2016b). Optimal Control Trajectories Given the above-defined model of neural dynamics, as well as the structural network representation extracted from diffusion imaging data, we can now formally address the question of how the activity in individual brain regions affects the trajectory of the brain as it transitions between states. We begin by defining an optimization problem to identify the trajectory between a specified pair of brain states that minimizes a given cost function. We define a cost function by the weighted sum of the energy cost of the transition and the integrated squared distance between the transition states and the target state. We choose this dual-term cost function for two reasons. First, theoretically, the energy cost term constrains the range of the time-dependent control energy u(t). In practice, this means that the brain cannot use an infinite amount of energy to perform the task (i.e., elicit the state transition), a constraint that is consistent with the natural energetic restrictions implicit in the nature of all biological systems but particularly neural systems (Niven & Laughlin, 2008; Laughlin et al., 1998; Attwell & Laughlin, 2001; Laughlin, 2001). Second, the term of the integrated distance term provides a direct constraint on the trajectory. Mathematically, this constraint penalizes trajectories that traverse states that are far away from the target state, based on the intuition that optimal transitions between states should possess reasonable lengths rather than being characterized by a random walk in state space. Together, these two terms in the cost function enable us to define an optimal control model from which we expect to find trajectories (from a given initial state to a specified target state) characterized by a balance between energy cost and trajectory length. In the context of the optimization problem defined above, we wish to determine the trajectory from an initial state x0 to a target state xT . To do so, it suffices to solve the variational problem with the constraints from Equation 1 and the boundary conditions for x(t), i.e. x(0) is the initial state and x(T ) is the target state. Note that here, the variational problem does not refer to the Bayesian variational inference, which tries to approximate an intractable posterior distribution. Instead, we use the term in the more traditional sense, and address the variational problem to infer a control input function u(t) to minimize the cost functional defined in Equation [4] with the boundary constraints. Mathematically, the variational problem is formulated as (cid:0)(xT − x(t))T (xT − x(t)) + ρu(t)T u(t)(cid:1) dt, (cid:90) T min u s.t. 0 x(t) = Ax(t) + Bu(t), x(0) = x0, x(T ) = xT , where T is the control horizon, ρ ∈ R>0, and (xT − x(t)) is the distance between the state at time t and the target state. To compute an optimal control u∗ that induces a transition from the initial state x0 to the target state xT , we define the Hamiltonian as (4) From the Pontryagin minimum principle (Boltyanskii et al., 1960), if u∗ is an optimal solution to the minimization problem with corresponding state trajectory x∗, then there exists p∗ such that H(p, x, u, t) = xT x + ρuT u + pT (Ax + Bu). which reduces to Next, we denote ) + AT p∗ = − p∗, ∂H ∂x ∂H ∂u (cid:21) (cid:20) x∗ p∗ = −2(xT − x∗ = 2ρu∗ + BT p∗ (cid:20) A −(2ρ)−1BBT = −2I 2xT (cid:20)0 (cid:21) I , + (cid:21) = 0. p∗ −AT (cid:21) (cid:21)(cid:20)x∗ (cid:20) A −(2ρ)−1BBT (cid:20)x∗ (cid:21) (cid:21) (cid:20)0 −AT −2I p∗ , 2xT , I A = x = b = then Eqn [7] can be written as x = Ax + b, 4 (3) (5) (6) (7) (8) (9) (10) (11) from which we can derive that x + A−1 b = eAtc, where c is a constant to be fixed from the boundary conditions. Let t = 0, T with the corresponding x0 and xT , we have (cid:35) (cid:34) b1 b2 = A−1 b, e−AT = (cid:20)E11 E12 (cid:21) E21 E22 (12) and plug in (cid:20)x(0) (cid:20)x(T ) p(0) (cid:21) (cid:21) p(T ) + + (cid:35) (cid:35) (cid:34) b1 (cid:34) b1 b2 b2 = = b = (cid:20) c1 (cid:21) (cid:20)E11 E12 c2 , (cid:21)−1(cid:20) c1 (cid:21) . E21 E22 c2 Note that from Equation[13], we can solve for c1, where c1 = x(0) + b1. Finally, with c1 on hand from Eqn[14], we can compute p(T ), where b1 − E12 p(T ) = E−1 12 (c1 − E11 b2 − E11x(T )), with which we can finally get c2, where c2 = E21x(T ) + E22p(T ) + E21 b1 + E22 b2 (13) (14) (15) (16) (17) (20) (21) (cid:90) T (cid:90) T 0 and further the u(t) and x(t) from Equation 12. Note that the formulae we derive here are the closed form solutions to the optimization objective, and therefore a numerical solver is not needed. Statistics of Optimal Control Trajectories After calculating the optimal trajectories between initial and final states, we next sought to address the question of whether these trajectories differed in their energetic and spatial requirements for different choices of control strategies, and between individual's whose brains were healthy and normally functioning, and individuals who had experienced a mild traumatic brain injury and had presented with complaints of mild cognitive impairment. To address this question, we computed the energy cost of a trajectory, integrated over time T , as E(K, x0, xT ) = u2K,x0,xT dt, (18) and the spatial cost of a trajectory, integrated over time T , as (19) where uK,x0,xT is the associated control input and xK,x0,xT is the controlled trajectory with the given control set K, initial state x0 and the target state xT . We treat this energy as a simple statistic that can be compared across trajectories and subject groups, as an indirect measure from which we may infer optimality of cognitive function. 0 x2K,x0,xT dt, S(K, x0, xT ) = Control Efficiency The control efficiency is defined for each region to quantify its efficiency in affecting the transition from the default mode state to the three target states. Mathematically, suppose we have N randomly chosen control sets, each indexed by K1, . . . ,KN , for the target states xj T , j = 1, 2, 3, we calculate the corresponding optimal trajectory with respect to Kk and denote the energy cost of the trajectory as E(Kk, x0, xj T is then defined as T ). The tiered value of the control set Kk for target xj N(cid:88) l=1 tkj = 1(E(Kl, x0, xj T ) > E(Kk, x0, xj T )) where lower energy costs imply higher tiered values. The control efficiency for node i in task j is then or intuitively, the average of these tiered values. ζij = (cid:80)N (cid:80)N k=1 1(i ∈ Kk) · tkj k=1 1(i ∈ Kk) . 5 Network Communicability to the Target State G = exp(D−1/2AD−1/2), where D is the diagonal matrix with the diagonal element Dii =(cid:80) communicability to all of the target regions, i.e. GTi =(cid:80) For a given weighted network A, the network communicability G quantifies the extent of indirect connectivity among nodes. Here we adopt the generalized definition in (Crofts & Higham, 2009) and define the network communicability as i Aij. For a given target state xT , denote the set of active regions as IxT , the communicability to the target states (GTi) is then defined as the sum of Gij. Further, the normalized network communicability to the target regions (Ci) is then defined as . (22) j∈IxT GTi(cid:80) j GTj Ci = All results reported in this study are based on the normalized network communicability. Energetic Impact of Brain Regions on Control Trajectories To quantify the robustness of controllability of a node when it is removed from the control set consisting of all nodes, we iteratively remove nodes from the network and compute the energetic impact of each region on the optimal trajectory as the resulting increase in the log value of the energy cost. Intuitively, regions with high energetic impact are those whose removal from the network causes the greatest increase in the energy required for the state transition. Mathematically, denote K0 as the control set of all nodes and Ki as the control set without node [K]i, the energetic impact of node i for target xj T is defined as Iij = log E(Ki, x0, xj T ) E(K0, x0, xj T ) (23) which intuitively measures robustness controllability. Results To begin, we set the initial state of the brain to be an activation pattern consistent with those empirically observed in the brain's baseline condition. More specifically, we set the initial state such that the regions of the default mode network had activity magnitudes equal to 1 ("on"), while all other regions had activity magnitudes equal to 0 ("off"). Furthermore, we examined 3 distinct target states such that regions of the (i) auditory, (ii) extended visual, or (iii) motor systems had activity magnitudes equal to 1 ("on"), while all other regions had activity magnitudes equal to 0 ("off"). In this context, we sought to understand characteristics of the transitions between initial and target states that could be performed with minimal energy, minimal time, and along short trajectories in state space by multiple control regions (multi-point control; see Fig. 1C and Methods). We note that mathematically, we measure time in arbitrary units, at each of which control energy can be utilized by a brain region. Intuitively, we operationalize time as consistent with the temporal scale at which brain regions can alter their activity magnitudes to affect state transitions. Characteristics of Optimal Control Trajectories We first study the three state transitions from the default mode to (i) auditory, (ii) extended visual, and (iii) motor states (Fig. 2A). We take a hypothesis-driven approach and define the "control set" to be composed of dorsal and ventral attention (Posner & Petersen, 1989), fronto-parietal, and cingulo-opercular cognitive control regions (Gu et al., 2015). That is, this set of 87 regions will utilize control energy using a multi-point control strategy, thereby changing the time-varying activity magnitudes of all brain regions (Fig. 2B). The optimal trajectories display multiple peaks in the distance from the target state as a function of time, and are altered very little by whether the target state is the auditory, extended visual, or motor system (Fig. 2C). Because the optimal trajectory is determined via a balance of control energy and trajectory distance (see Methods), it stands to reason that the time-dependent energy utilized by the control set is inversely related to the distance between the current state and the target state. When little control energy is utilized, the current state can drift far from the target state, while when a larger magnitude of control energy is utilized, the current state moves closer to the target state (Fig. 2D). It is important to note that these general characteristics of the optimal control trajectories are dependent on our choice of the control set (which here we guide with biologically motivated hypotheses), as well as on a penalty on the time required for the transition (ρ in Equation[3]; see Methods). In the supplement, we examine the effect of alternative choices for both the control set and ρ. First, we find that when the control set includes every node in the network, the distance to the target state decreases monotonically to zero along the trajectory (Fig. S1A). Second, we consider the effect of the penalty term on control energy, ρ. For the results presented here, we fix ρ to be equal to 1. However, in the supplement, we explore a wide range of ρ values, and show that when ρ is small, the optimal control trajectory is largely driven by a minimization of the integrated squared distance to the target. In contrast, when ρ is large, the optimal control trajectory is largely driven by the magnitude of the utilized energy (Fig. S1B). Importantly, we did not perform a full sweep of ρ from 0 to infinity because very small values of ρ cause numeric instabilities in the calculations. 6 Figure 2: Optimal Control Trajectories. (A) We study 3 distinct types of state transitions in which the initial state is characterized by high activity in the default mode system, and the target states are characterized by high activity in auditory (blue), extended visual (green), or motor (red) systems. (B) The activation profiles of all N = 234 brain regions as a function of time along the optimal control trajectory, illustrating that activity magnitudes vary by region and by time. Activation can be either positive or negative and the exact range of values will depend on the initial state, the target state, and the control set. Regions are listed in the following order: initial state, target state, controllers and others. (C) The average distance from the current state x(t) to the target state x(T ) as a function of time for the trajectories from the default mode system to the auditory, visual, and motor systems, illustrating behavior in the large state space. (D) The average control energy utilized by the control set as a function of time for the trajectories from the default mode system to the auditory, visual, and motor systems. The similarity of the curves observed in panels (C) and (D) is driven largely by the fact that they share the same control set. See Fig. S2(B) for additional information on the range of these control energy values along the trajectories. Colors representing target states are identical in panels (A), (C), and (D). 7 BACDDefault ModeVisual ModeSomatosensory ModeAuditory ModeEnergyEnergyEnergyTime (a.u.)2004006008001000Region50100150200-10010Time (a.u.)2004006008001000Region50100150200-10010Time (a.u.)2004006008001000Region50100150200-10010Time (a.u.)05001000Trajectory Distance, xt-xf0102030405060Time (a.u.)05001000Control Energy, u0200400600800State Value, xState Value, xState Value, x Structurally-Driven Task Preference for Control Regions We next ask whether certain brain regions are located at specific points in the structural network that make them predisposed to play consistent and important roles in driving optimal control trajectories. To answer this question, we choose control sets of the same size as the brain's hypothesized cognitive control set; recall that in the previous section, we defined the brain's cognitive control set to consist of the 87 nodes of the dorsal and ventral attention, fronto-parietal, and cingulo- opercular systems following (Gu et al., 2015). Here, we choose the 87 regions of these new control sets uniformly at random from the set of all nodes. Using these "random" control sets, we computed the optimal control trajectory for each of the three state transitions and for each subject separately. Then, we rank the random control sets in descending order according to the energy cost of the trajectory and we assign every region participating in an r-ranked control set with rank-value r. Next, we define the control efficiency of a brain region to be the sum of its rank values in all of the random control sets it belongs divided by the total number of sets it belongs to. Intuitively, a region with a high control efficiency is one that exerts control with little energy utilization. Importantly, it must decrease activation in the initial state, and increase activation in the target state, a pair of capabilities that depend on the pattern of connections emanating from the region. In general, we observe that a region's preference for being an optimal controller (exerting control with little energy utilization) is positively correlated with its network communicability to the regions of high activity in the target state (Spearman correlation r = 0.27, p < 4.8 × 10−4; see Fig. 3A). We recall that network communicability is a measurement of the strength of a connection from one region to another that accounts for walks of all lengths (see Methods). Interestingly, we observed this same correlation between control efficiency and network communicability across optimal control trajectories for all three state transitions, from the default mode to the auditory (r = 0.36, p = 1.4 × 10−8), extended visual (r = 0.51, p = 1.1× 10−16), or motor (r = 0.42, p = 2.1× 10−11) systems (Fig. 3B-D). Together, these results indicate that regions that are close (in terms of walk lengths) to regions of high activity in the target state are efficient controllers for that specific state transition. Note that these regions are not purely target areas, likely due to the fact that they must also decrease activation in the initial state. The general role that network proximity to the target state plays for control regions ensures that regions that are proximate to all three target states (auditory, extended visual, and motor) will be consistent controllers, while regions that are proximate to only one of the target states will be task-specific controllers. To better understand the anatomy of efficient controllers, we transformed control efficiency values to z-scores and defined an efficient control hub to be any region whose associated p-value was less than 0.025. Across all three state transitions, we found that the supramarginal gyrus specifically, and the inferior parietal lobule more generally, consistently acted as efficient control hubs. The consistent control role of these regions is likely due to the fact that these areas are structurally interconnected with ventral premotor cortex, a key input to primary sensorimotor areas (Kandel et al., 2000). The areas that are more specific to the three state transitions include medial parietal cortex (motor transition), orbitofrontal and inferior temporal cortex (visual transition), and superior temporal cortex (auditory transition). Regional Roles in Control Tasks The analyses outlined above are built on the assumption that the brain uses fronto-parietal, cingulo-opercular, and attention systems to affect cognitive control, which we define as the ability to move the brain from an initial state (e.g., the default mode systems) to a specified final state (e.g., activation of extended visual, auditory, or motor cortex). However, one might naturally ask whether these regions of the brain could have been predicted a priori to be effective controllers based on traditional engineering-based notions of control. In the control theory literature, particularly the literature devoted to the subfield of network controllability, there exist several controllability notions, including average, modal, and boundary control (Pasqualetti et al., 2014). Average controllability identifies brain areas that can theoretically steer the system into many different states, or patterns of neurophysiological activity magnitudes across brain regions. Modal controllability identifies brain areas that can theoretically steer the system into difficult-to-reach states. Boundary controllability identifies brain areas that can theoretically steer the system into states where different cognitive systems are either coupled or decoupled. See the SI for mathematical definitions and (Gu et al., 2015) for prior studies in human neuroimaging. We calculated average, modal, and boundary control values for each node in the network. We observe that while cognitive control regions cover a broad swath of frontal and parietal cortex, including medial frontal cortex and anterior cingulate (Fig. 4A), the number of these regions that intersect with the strongest 87 average, modal, or boundary control hubs was on average approximately 50 (Fig. 4B). These results suggest that the control capabilities of the human brain's cognitive control regions may not be perfectly aligned with control notions previously developed in the field of mechanical engineering, provided that the model assumptions and data quality are appropriate (see Methodological Considerations). Instead, cognitive control regions in the human brain may have distinct capabilities necessary for the specific transitions required by the brain under the constraints imposed by neuroanatomy and neurophysiology. To more directly test this possibility, we examined the average distance (Fig. 4C) and energy (Fig. 4D) for transitions from the default mode to the auditory, extended visual, and sensorimotor states that are driven by average, modal, and boundary control hubs, or by regions of fronto-parietal, cingulo-opercular, and attention systems. We observed that both the trajectory cost and the energy cost differ by control strategy and by target state. We quantify this observation using a 2-way ANOVA with both the control strategy and target state as categorical factors. Using the trajectory cost as the dependent variable, we observed a significant main effect of control strategy (F = 78.74, p = 4.65× 10−41), a significant main 8 Figure 3: Structurally-Driven Task Preference for Control Regions. (A) Top Regions with high control efficiency (see Eqn 21) across all 3 state transitions: from the default mode to auditory, extended visual, and motor systems. Bottom Scatterplot of the control efficiency with the average network communicability to all 3 target regions (Spearman correlation r = 0.27, p < 4.8 × 10−4). (B -- D) Top Regions with high control efficiency for the transition from default mode to (B) motor, (C) extended visual, and (D) auditory (r = 0.36, p = 1.4 × 10−8) targets (top). Bottom Scatter plot of control efficiency versus normalized network communicability with regions that are active in the target state: motor (r = 0.42, p = 2.1 × 10−11), extended visual (r = 0.51, p = 1.1 × 10−16), and auditory (r = 0.36, p = 1.4 × 10−8). Values of control efficiency in all four panels are averaged of subjects. effect of target state (F = 29.24, p = 1.12 × 10−12), and a significant interaction between control strategy and target state (F = 11.36, p = 7.6× 10−12). Similarly, using the energy cost as the dependent variable, we observed a significant main effect of control strategy (F = 67.94, p = 2.48 × 10−36), a significant main effect of target state (F = 39.18, p = 1.99 × 10−16), and a significant interaction between control strategy and target state (F = 10.93, p = 2.18 × 10−11). Collapsing over target states and performing post-hoc testing, we observed that cognitive control regions displayed a similar average trajectory cost to average control hubs, but a lower average trajectory cost than modal and boundary control hubs (p < 0.05 uncorrected). Furthermore, cognitive control regions possessed a higher average energy cost than the average and modal control hubs, but a lower average energy cost than the boundary control hubs. These results interestingly suggest that the human's cognitive control regions, as defined by decades of research in cognitive neuroscience, may affect state transitions using neither the shortest distances nor the lowest energies possible, provided that the model assumptions and data quality are appropriate (see Methodological Considerations). This is likely due to the fact that cognitive control regions must affect a broad array of state transitions that cannot easily be classified into average, modal, and boundary control strategies. Specificity of Control in Health and Following Injury The unique role of brain regions in affecting control strategies may bring with it vulnerability to injury. When a brain network is injured, regional control roles may be significantly altered, potentially increasing susceptibility to underlying abnormalities in neuronal dynamics. To characterize this vulnerability, we determine the degree to which a single brain region impacts putative control processes and we ask whether that specificity is maintained or altered following brain injury. We measure specificity by iteratively removing nodes from the control set, and we compute the energetic impact of each region on the optimal trajectory as the resulting increase in the log value of the energy cost (see Fig. 5A and Eqn 23 in Methods). Intuitively, regions with high energetic impact are those whose removal from the network causes the greatest increase in the energy required for the state transition. Across all subjects and all tasks, we observe that the regions with the highest energy impact are the supramarginal gyrus specifically, and the inferior parietal lobule more generally, the same regions that emerged as consistent and efficient controllers in Fig. 2A. Next we determined whether energetic impact -- our proxy for regional specificity of control roles -- is altered in individuals with mild traumatic brain injury (mTBI). Intuitively, if all regions of a brain have high energetic impact, this indicates that each region is performing a different control role which is destroyed by removal of the node. By contrast, if all regions of a brain have low average energetic impact, this indicates that each region is performing a similar control role that is not destroyed by removal of a node. We observed that individuals with mTBI displayed anatomically similar patterns of energetic impact on control trajectories as regions are removed from the network (Fig. 5B). However, the average magnitude and variability of the energetic impact differed significantly between the two groups, with individuals having experienced mTBI displaying 9 SomatosensoryControl E(cid:31)ciencyControl E(cid:31)ciency Control E(cid:31)ciencty Control E(cid:31)ciencyRank of Communicability with TargetsRank of Communicability with TargetsRank of Communicability with TargetsRank of Communicability with TargetsVisualABCDCombinationAuditory050100150200250×10-33.84.04.24.44.64.85050100150200250×10-33.844.24.44.64.85050100150200250×10-33.844.24.44.64.85050100150200250×10-33.844.24.44.64.85 Figure 4: Regional Roles in Control Tasks. (A) Cognitive control regions cover a broad swath of frontal and parietal cortex, including medial frontal cortex and anterior cingulate, and are defined as regions included in fronto-parietal, cingulo-opercular, and attention systems (Gu et al., 2015). (B) The number of these regions overlapping with the strongest 87 average, modal and boundary control hubs is approximately 50. Different choices of control strategies result in variation in both (C) trajectory cost and (D) energy cost. Here, HC refers to cognitive control regions, AC refers to average control hubs, MC refers to modal control hubs, and BC refers to boundary control hubs. significantly lower values of average magnitude of energetic impact (permutation test: p = 5.0 × 10−6) and lower values of the average standard deviation of energetic impact (p = 2.0 × 10−6). We note that common graphic metrics including the degree, path length, clustering coefficient, modularity, local efficiency, global efficiency, and density were not significantly different between the two groups, suggesting that this effect is specific to control (see Supplement). These results indicate that mTBI patients display a loss of specificity in the putative control roles of brain regions, suggesting greater susceptibility to damage-induced noise in neurophysiological processes, or to external drivers in the form of stimulation. Discussion Here we ask whether structural connectivity forms a fundamental constraint on how the brain may move between diverse cognitive states. To address this question, we capitalize on recent advances in network control theory to identify and characterize optimal trajectories from an initial state (composed of high activity in the default mode system) to target states (composed of high activity in sensorimotor systems) with finite time, limited energy, and multi-point control. Using structural brain networks estimated from diffusion imaging data acquired in a large cohort of 48 healthy individuals and 11 patients with mild traumatic brain injury, we show that these optimal control trajectories are characterized by continuous changes in regional activity across the brain. We show that the regions critical for eliciting these state transitions differ depending on the target state, but that heteromodal association hubs -- predominantly in the supramarginal gyrus specifically, and the inferior parietal lobule more generally -- are consistently recruited for all three transitions. Finally, we study the sensitivity of optimal control trajectories to the removal of nodes from the network, and we demonstrate that brain networks from individuals with mTBI display maladaptive control capabilities suggestive of a limited dynamic range of states available to the system. Together, these results offer initial insights into how structural network differences between individuals impact their potential to control transitions between cognitive states. Role of Structural Connectivity in Shaping Brain Functional Patterns A growing body of literature on the relationship between brain structure and function has demonstrated that the brain's network of anatomical connections constrains the range of spontaneous (Deco et al., 2011) and task-related (Hermundstad et al., 2013b) fluctuations in brain activity. Evidence for such structural underpinnings comes from two distinct lines of research. On one hand, empirical studies have demonstrated that structural insults in the form of lesions result in acute reorganization of the brain's pattern of functional coupling (Johnston et al., 2008; O'Reilly et al., 2013). These observations are further buttressed by simulation studies in which structural connectivity has been used to constrain interactions among dynamic elements in biophysical models of brain activity (Honey et al., 2007, 2009; Adachi et al., 2011) and models of network communication (Goni et al., 2014; Abdelnour et al., 2014; Misi´c et al., 2015). Though this forward modeling approach has 10 Control StrategiesControl StrategiesControl StrategiesControl StrategiesControl StrategiesControl StrategiesControl StrategiesNumber of Common NoddesFronto-parietal, cingulo-opercular,and attention systemsDMN --> SMDMN --> VSDMN --> ADDMN --> ADDMN --> VSDMN --> SMABCD253035404550ACACMCBCHCACMCBCHCACMCBCHCACMCBCHCACMCBCHCACMCBCHCMCBC468101214567891011121346810121416186810121416182010111213141516171881012141618202224Log Energy CostLog Energy CostLog Energy CostLog Spatial CostLog Spatial CostLog Spatial Cost01 Figure 5: Specificity of Control in Health and Following Injury. (A) Theoretically, the brain is fully controllable when every region is a control point, but may not be fully controllable when fewer regions are used to affect control. (B) The regions with the highest values of energetic impact on control trajectories upon removal from the network, on average across subjects and tasks, were the supramarginal gyrus specifically, and the inferior parietal lobule more generally. In general, the healthy group and the mTBI group displayed similar anatomical patterns of energetic impact. (C) Magnitude and standard derivation of energetic impact averaged over regions and tasks; boxplots indicate variation over subjects. Even after removing the single outlier in the healthy group, patients with mTBI displayed significantly lower values of average magnitude of energetic impact (permutation test: p = 1.1 × 10−5) and lower values of the average standard deviation of energetic impact (p = 2.0 × 10−6) than healthy controls. 11 Controlled RegionsControlled RegionsControlled RegionsUncontrolled RegionsThe Fully Controlled BrainThe Almost Fully Controlled BrainInitial StateTarget StateHealthy GroupHealthy GroupTBI GroupHealthy GroupTBI GroupE0E1∆EABDCMean of Energetic ImpactStandard Deviation of Energetic Impact0.0017mTBI Group0.68910.00170.7180Energetic ImpactEnergetic Impact0.120.140.160.180.20.22p = 5.0 x 10-60.30.350.40.450.50.55p = 2.0 x 10-6 proven fruitful in predicting observed patterns of functional connectivity, the precise mapping of brain structure to function remains unclear. The present study builds on this body of work, using a dynamical model of how brain activity propagates over a network in order to gain insight into what features of that network facilitate easy transitions from a baseline (default mode) state to states where the brain's primary sensorimotor systems are activated. In contrast to previous simulation studies that have focused on network features that influence the passive spread of activity over time, this present study directly engages the question of how those same features enable the state of the system to be controlled. We use this model to demonstrate that brain regions are differentially-suited for particular control tasks, roles that can be predicted on the basis of how well-connected they are to regions in the target state. Regions that are close (in terms of walk lengths) to regions of high activity in the target state are efficient controllers for that specific state transition. It follows, then, that a brain region's capacity to dynamically influence a network depends not only on its pattern of connectivity, but also the repertoire of states that the system visits. In other words, a region that maintains many connections (both direct and indirect), but never to regions that are "active" in target states, may exert less influence than a region that maintains few connections, but whose connections are distributed among regions that are "active" in many target states. We further demonstrate that this mapping of brain structure to specific functions is altered in individuals with mTBI, suggesting that injury may alter control profiles of individual brain regions. Our finding that the inferior parietal lobule forms a consistently effective control region, across all three target states, is particularly interesting when considered in the context of prior literature on this region's structural and functional roles. In particular, the inferior parietal lobule represents the superior portion of the temporoparietal junction, a multimodal area associated with functions as wide ranging as calculation, finger gnosis, left/right orientation, and writing (Rusconi et al., 2009). Focal damage to this area leads to wide-spread cognitive disfunction (e.g., Gerstmann syndrome) as a consequence of the unique confluence of white matter pathways underlying this region (Rushworth et al., 2006). The diverse white matter projections emanating from this area may support its putative role in effectively controlling brain function. Indeed, recent evidence suggests that the right temporoparietal junction links two antagonistic brain networks processing external versus internal information: a midcingulate-motor-insular network associated with attention, and a parietal network associated with social cognition and memory retrieval (Bzdok et al., 2013). These data support the notion that the right temporoparietal junction controls our attention to salient external events (Corbetta & Shulman, 2002), perhaps with early input from the right fronto-insular cortex thought to drive switching between central-executive and default-mode networks (Sridharan et al., 2008). Single versus Multipoint Control An important feature of our model lies in the delineation of a control set, a group of brain regions that can affect distributed control. The focus on multiple points of control throughout the system is one that has important theoretical motivations and empirical correlates. Prior computational models demonstrate that while the brain is theoretically controllable via input to a single control point, the energy and time required for that control is such that the brain is practically uncontrollable (Gu et al., 2015). These data argue for an assessment of multi-point control as a better proxy of control strategies that the brain might utilize. Indeed, such an argument is consistent with empirical observations that stimulation (or even drug manipulations) focused on single brain regions are less effective in treating psychiatric disease than interventions that target multiple brain regions (Sommer et al., 2012; Tortella et al., 2014). A prime empirical example of multi-point control is cognitive behavioral therapy, which offers a spatio-temporal pattern of activations that enhances cognitive function and decreases psychiatric symptoms across diagnostic categories (Lett et al., 2014; Radhu et al., 2012; Cima et al., 2014). Other potential multi-point control mechanisms include grid stimulation across multiple electrodes, suggested in the control of medically refractory epilepsy (Ching et al., 2012). Diversity in Human Brain Control Strategies By studying multi-point control, we were able to directly assess whether the optimal control trajectories elicited from fronto-parietal, cingulo-opercular, and attention systems displayed similar distance and energy to trajectories to those ob- tained using regions selected for engineering-based notions of control (Pasqualetti et al., 2014). Specifically, we compare and contrast the performance of human cognitive control regions to average, modal, and boundary controllers Gu et al. (2015). One might naturally ask whether and how each of these engineering-based notions of controllers is an appropriate theoretical quantity to consider in the context of neurobiology. Average controllers are those theoretically able to push the system from any arbitrary initial state to any easily-reachable state, nearby on the energy landscape. Modal controllers are those that are optimally placed to move the system from any arbitrary initial state to any difficult-to-reach state, far away on the energy landscape. Boundary controllers are those that are optimally placed to integrate or segregate network communities in the system. Common assumptions that underlie each of these control strategies are that (i) controllers can be identified independently from the initial and final states, and (ii) all states in the energy landscape are accessible to the system. The common constraint on each of these strategies is that expended energy must be minimized for a region to be referred to as a controller. In the human brain, it is not well-known whether these assumptions are met, or whether the energetic constraint is sufficient to predict control functions. Interestingly, we observe that the optimal trajectories elicited by canonically defined cognitive control regions do not show similar energy requirements or trajectory distances to any of these previously described control types. There are several 12 potential reasons for this observation: (i) false negatives in the diffusion imaging data impacting on the observed network profiles, (ii) assumptions of the linear model, and (iii) bona fide differences between mechanical controllers and biological controllers. Prior work demonstrating robustness of controllability profiles across large cohorts and different diffusion imaging acquisition protocols provide initial evidence that the first explanation is unlikely to fully explain our findings (Gu et al., 2015). Regarding the second explanation -- assumptions of the linear model -- it is interesting to note that recent evidence suggests that the average, modal, and boundary controllability profiles identified by the linear model provide excellent predictions for the behavior of nonlinear models (Muldoon et al., 2016b). Evidence supporting the third interpretation -- that these observed differences are bona fide differences between mechanical controllers and biological controllers -- is provided by the fact that cognitive control regions must affect a broad array of state transitions that do not easily fit into prior classifications. These transitions include switching behavior (Hansen et al., 2015), inter-state competition (Cocchi et al., 2013), distributed rather than centralized control (Eisenreich et al., 2016), and push-pull control (Khambhati et al., 2016), which may each offer differential advantages for neural computations (Durstewitz & Deco, 2008). Maladaptive Control in Traumatic Brain Injury Finally, our assessment of patients with mild traumatic brain injury enabled us to determine whether widespread injury leads to a decrement in the healthy network control profiles, thus theoretically requiring greater energy for the same functions, or an enhancement of the healthy network control profiles at the cost of expected sensitivity to external perturbations. Our data provide initial evidence for maladaptive control of the latter sort in patients with mild traumatic brain injury. Understanding the impact of brain injury on cognitive processes, including the ability to switch between cognitive states, is a major goal in clinical neuroscience. Indeed, traumatic brain injury is a common source of brain dysfunction, affecting more than 200,000 individuals per year in the United States alone. Injuries -- often caused by motor vehicle and sports accidents -- result in damage to neuronal axons, including long-distance white matter fiber bundles (Johnson et al., 2013) as well as u-fibers and deep white matter tracks with multiple crossings. The pattern of injury can be multi-focal and variable across individuals (Kinnunen et al., 2011; Sidaros et al., 2008; Hellyer et al., 2013), challenging comprehensive predictors and generalizable interventions. Recent evidence suggests that injury-induced, widespread damage to white matter tracts critically impacts large-scale network organization in the human brain, as measured by diffusion imaging tractography (Kinnunen et al., 2011; Fagerholm et al., 2015). Moreover, this damage is associated with fundamental changes in cognitive function (Sharp et al., 2014), including information processing speed, executive function, and associative memory (Fagerholm et al., 2015). Each of these cognitive deficits intuitively depends on the ability to transition from one cognitive state to another; yet an understanding of structural drivers of these transitions and their potential alteration in mTBI has remained elusive. Here we demonstrate a loss of specificity in putative control processes in mTBI, suggesting that the unique roles of individual brain regions in supporting cognitive state transitions are damaged. It is intuitively plausible that this decrement in regional specificity of control leads to broad changes in functional dynamics, particularly in the system's susceptibility to damage-induced noise in neurophysiological processes (Garrett et al., 2013). Indeed, the observed decrements in energetic impact might further provide a direct structural mechanism for the decreased signal variability observed in mTBI using electrophysiological imaging (Raja Beharelle et al., 2012; Nenadovic et al., 2008). More generally, these findings highlight the fact that the healthy brain might display a degree of controllability that is either decremented or enhanced in injury and disease, suggesting the possibility of a U-shaped curve reminiscent of similar curves observed in other brain network phenotypes (Collin & van den Heuvel, 2013; Cools & D'Esposito, 2011). Methodological Considerations A few methodological points are worthy of additional consideration. First, in this study we examined structural brain networks derived from diffusion imaging data and associated tractography algorithms. These algorithms remain in their relative infancy, and can still report spurious tracts or fail to report existing tracts (Thomas et al., 2014; Reveley et al., 2015; Pestilli et al., 2014). Despite the evolving nature of diffusion protocols and tractography algorithms, preliminary data provide initial evidence that consistent controllability profiles can be robustly observed across large cohorts and different diffusion imaging acquisition protocols (Gu et al., 2015). Formal validation in axonal tracing studies in monkeys and other mammals (Jbabdi et al., 2013) remains the gold standard for these types of data. However, it is important to note that initial work supports the notion that much of the structure present in DSI connectivity matrices recapitulates known projections observed in tract tracing studies of the macaque Hagmann et al. (2008). Second, following (Gu et al., 2015; Betzel et al., 2016; Muldoon et al., 2016b), we employ a linear dynamical model, consistent with prior empirical studies demonstrating their ability to predict features of resting state fMRI data (Gal´an, 2008; Honey et al., 2009). This choice is to some degree predicated on the well-developed theoretical and analytical results in the engineering and physics literatures examining the relationship between control and network topology (Liu et al., 2011; Muller & Schuppert, 2011; Yan et al., 2012). Moreover, it is plausible that even these results using simple linear models may offer important intuitions for controlling nonlinear models of brain function. Indeed, theoretical work over the last several decades has demonstrated the utility of describing non-linear systems in terms of a linear approximation in the neighborhood of the system's equilibrium points (Luenberger, 1979). Very recent evidence has extended these intuitions to neuroimaging data, demonstrating that the average, modal, and boundary controllability profiles identified by the linear model can be used 13 to predict the behavior of nonlinear models in the form of Wilson-Cowan oscillators, which are commonly used to understand the dynamics of cortical columns (Muldoon et al., 2016b). Future directions An interesting hypothesis generated by the current framework is that control capabilities may be altered dimensionally across traditionally separated diagnostic groups that display dysconnectivity in network hubs, as measured by regions of high eigenvector centrality. Such a hypothesis builds on the now seminal dysconnection hypothesis in schizophrenia (Stephan et al., 2009), and expands it to include an explicit dynamical control component. Indeed, mounting evidence suggests that the overload or failure of brain network hubs may be a common neurophysiological mechanism of a range of neurological disorders including Alzheimer's disease, multiple sclerosis, traumatic brain injury and epilepsy (Stam, 2014). Alterations in these hubs can also be used to predict the progression of psychiatric disorders such as schizophrenia (Collin et al., 2015). In both neurological and psychiatric disorders, these changes to network hubs may alter the control capabilities of the individual, challenging the normal executive functions required for daily living. It is also intuitively plausible that normal variation in hub architecture may play a role in individual differences in control capabilities in healthy individuals, impacting on the speed with which they transition between cognitive states. These topics will form important provender for future work. Acknowledgements: D.S.B., S.G., and R.F.B. acknowledge support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Research Laboratory and the Army Research Office through contract numbers W911NF-10-2- 0022 and W911NF-14-1-0679,the National Institute of Mental Health (2-R01-DC-009209-11), the National Institute of Child Health and Human Development (1R01HD086888-01), the Office of Naval Research, and the National Science Foundation (BCS-1441502, BCS-1430087, and CAREER PHY-1554488). S.T.G. and P.R.D. acknowledge support from a Head Health Challenge grant from General Electric and the National Football League. F.P. acknowledges support from the National Science Foundation award #BCS 1430280. The content is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies. Author Contributions S.G. performed the analysis. M.C. preprocessed the data. S.T.G. and P.R.D. acquired the data. S.G., R.F.B., F.P., D.S.B. developed the project. S.G., R.F.B., F.P., D.S.B. wrote the paper. Competing Interests The authors declare that they have no competing financial interests. References References Abdelnour, F., Voss, H. U., & Raj, A. (2014). Network diffusion accurately models the relationship between structural and functional brain connectivity networks. Neuroimage, 90 , 335 -- 347. Adachi, Y., Osada, T., Sporns, O., Watanabe, T., Matsui, T., Miyamoto, K., & Miyashita, Y. (2011). Functional connectivity between anatomically unconnected areas is shaped by collective network-level effects in the macaque cortex. Cerebral cortex , (p. bhr234). Alavash, M., Hilgetag, C. C., Thiel, C. M., & Giessing, C. (2015). Persistency and flexibility of complex brain networks underlie dual-task interference. Human brain mapping, 36 , 3542 -- 3562. Attwell, D., & Laughlin, S. B. (2001). An energy budget for signaling in the grey matter of the brain. J Cereb Blood Flow Metab, 21 , 1133 -- 1145. Bassett, D. S., Brown, J. A., Deshpande, V., Carlson, J. M., & Grafton, S. T. (2011a). Conserved and variable architecture of human white matter connectivity. Neuroimage, 54 , 1262 -- 1279. Bassett, D. S., Greenfield, D. L., Meyer-Lindenberg, A., Weinberger, D. R., Moore, S. W., & Bullmore, E. T. (2010). Efficient physical embedding of topologically complex information processing networks in brains and computer circuits. PLoS Comput Biol , 6 , e1000748. 14 Bassett, D. S., Wymbs, N. F., Porter, M. A., Mucha, P. J., Carlson, J. M., & Grafton, S. T. (2011b). Dynamic reconfiguration of human brain networks during learning. Proc Natl Acad Sci U S A, 108 , 7641 -- 7646. Bassett, D. S., Wymbs, N. F., Rombach, M. P., Porter, M. A., Mucha, P. J., & Grafton, S. T. (2013). Task-based core- periphery organization of human brain dynamics. PLoS Comput Biol , 9 , e1003171. Bassett, D. S., Yang, M., Wymbs, N. F., & Grafton, S. T. (2015). Learning-induced autonomy of sensorimotor systems. Nat Neurosci , 18 , 744 -- 751. Betzel, R. F., Gu, S., Medaglia, J. D., Pasqualetti, F., & Bassett, D. S. (2016). Optimally controlling the human connectome: the role of network topology. arXiv preprint arXiv:1603.05261 , . Boltyanskii, V. G., Gamkrelidze, R. V., & Pontryagin, L. S. (1960). The theory of optimal processes. I. The maximum principle. Technical Report DTIC Document. Braun, U., Muldoon, S. F., & Bassett, D. S. (2015). On human brain networks in health and disease. eLS , . Bullmore, E., & Sporns, O. (2009). Complex brain networks: graph theoretical analysis of structural and functional systems. Nature Reviews Neuroscience, 10 , 186 -- 198. Bzdok, D., Langner, R., Schilbach, L., Jakobs, O., Roski, C., Caspers, S., Laird, A. R., Fox, P. T., Zilles, K., & Eickhoff, S. B. (2013). Characterization of the temporo-parietal junction by combining data-driven parcellation, complementary connectivity analyses, and functional decoding. Neuroimage, 81 , 381 -- 392. Cammoun, L., Gigandet, X., Meskaldji, D., Thiran, J. P., Sporns, O., Do, K. Q., Maeder, P., Meuli, R., & Hagmann, P. (2012). Mapping the human connectome at multiple scales with diffusion spectrum MRI. J Neurosci Methods, 203 , 386 -- 397. Carter, A. R., Patel, K. R., Astafiev, S. V., Snyder, A. Z., Rengachary, J., Strube, M. J., Pope, A., Shimony, J. S., Lang, C. E., Shulman, G. L., & Corbetta, M. (2012). Upstream dysfunction of somatomotor functional connectivity after corticospinal damage in stroke. Neurorehabil Neural Repair , 26 , 7 -- 19. Ching, S., Brown, E. N., & Kramer, M. A. (2012). Distributed control in a mean-field cortical network model: implications for seizure suppression. Phys Rev E Stat Nonlin Soft Matter Phys, 86 , 021920. Cieslak, M., & Grafton, S. (2014). Local termination pattern analysis: a tool for comparing white matter morphology. Brain imaging and behavior , 8 , 292 -- 299. Cima, R. F., Andersson, G., Schmidt, C. J., & Henry, J. A. (2014). Cognitive-behavioral treatments for tinnitus: a review of the literature. J Am Acad Audiol , 25 , 29 -- 61. Cocchi, L., Zalesky, A., Fornito, A., & Mattingley, J. B. (2013). Dynamic cooperation and competition between brain systems during cognitive control. Trends in cognitive sciences, 17 , 493 -- 501. Collin, G., & van den Heuvel, M. P. (2013). The ontogeny of the human connectome: development and dynamic changes of brain connectivity across the life span. Neuroscientist, 19 , 616 -- 628. Collin, G., de Nijs, J., Hulshoff Pol, H. E., Cahn, W., & van den Heuvel, M. P. (2015). Connectome organization is related to longitudinal changes in general functioning, symptoms and IQ in chronic schizophrenia. Schizophr Res, S0920 -- 9964 , 00141 -- 00143. Cools, R., & D'Esposito, M. (2011). Inverted-U-shaped dopamine actions on human working memory and cognitive control. Biol Psychiatry, 69 , e113 -- e125. Corbetta, M., & Shulman, G. L. (2002). Control of goal-directed and stimulus-driven attention in the brain. Nat Rev Neurosci , 3 , 201 -- 215. Crofts, J. J., & Higham, D. J. (2009). A weighted communicability measure applied to complex brain networks. Journal of the Royal Society Interface, (pp. rsif -- 2008). Daducci, A., Gerhard, S., Griffa, A., Lemkaddem, A., Cammoun, L., Gigandet, X., Meuli, R., Hagmann, P., & Thiran, J. P. (2012). The connectome mapper: an open-source processing pipeline to map connectomes with MRI. PLoS One, 7 , e48121. Deco, G., Jirsa, V. K., & McIntosh, A. R. (2011). Emerging concepts for the dynamical organization of resting-state activity in the brain. Nat Rev Neurosci , 12 , 43 -- 56. 15 Di Martino, A., Fair, D. A., Kelly, C., Satterthwaite, T. D., Castellanos, F. X., Thomason, M. E., Craddock, R. C., Luna, B., Leventhal, B. L., Zuo, X.-N. et al. (2014). Unraveling the miswired connectome: a developmental perspective. Neuron, 83 , 1335 -- 1353. Durstewitz, D., & Deco, G. (2008). Computational significance of transient dynamics in cortical networks. Eur J Neurosci , 27 , 217 -- 227. Eisenreich, B., Akaishi, R., & Hayden, B. (2016). Control without controllers: Towards a distributed neuroscience of executive control. bioRxiv , 1101 , 077685. Fagerholm, E. D., Hellyer, P. J., Scott, G., Leech, R., & Sharp, D. J. (2015). Disconnection of network hubs and cognitive impairment after traumatic brain injury. Brain, 138 , 1696 -- 1709. Fiete, I. R., Senn, W., Wang, C. Z., & Hahnloser, R. H. R. (2010). Spike-time-dependent plasticity and heterosynaptic competition organize networks to produce long scale-free sequences of neural activity. Neuron, 65 , 563 -- 576. Freeman, W. J. (1994). Characterization of state transitions in spatially distributed, chaotic, nonlinear, dynamical systems in cerebral cortex. Integr Physiol Behav Sci , 29 , 294 -- 306. Gal´an, R. F. (2008). On how network architecture determines the dominant patterns of spontaneous neural activity. PLoS One, 3 , e2148. Garrett, D. D., Samanez-Larkin, G. R., MacDonald, S. W., Lindenberger, U., McIntosh, A. R., & Grady, C. L. (2013). Moment-to-moment brain signal variability: a next frontier in human brain mapping? Neurosci Biobehav Rev , 37 , 610 -- 624. Gazzaniga, M. S. (Ed.) (2013). The cognitive neurosciences. MIT Press. Goni, J., van den Heuvel, M. P., Avena-Koenigsberger, A., de Mendizabal, N. V., Betzel, R. F., Griffa, A., Hagmann, P., Corominas-Murtra, B., Thiran, J.-P., & Sporns, O. (2014). Resting-brain functional connectivity predicted by analytic measures of network communication. Proceedings of the National Academy of Sciences, 111 , 833 -- 838. Gu, S., Cieslak, M., Baird, B., Muldoon, S. F., Grafton, S. T., Pasqualetti, F., & Bassett, D. S. (2016). The energy landscape of neurophysiological activity implicit in brain network structure. Submitted , . Gu, S., Pasqualetti, F., Cieslak, M., Telesford, Q. K., Alfred, B. Y., Kahn, A. E., Medaglia, J. D., Vettel, J. M., Miller, M. B., Grafton, S. T. et al. (2015). Controllability of structural brain networks. Nature communications, 6 . Hagmann, P., Cammoun, L., Gigandet, X., Meuli, R., Honey, C. J., Wedeen, V. J., & Sporns, O. (2008). Mapping the structural core of human cerebral cortex. PLoS Biol , 6 , e159. Hansen, E. C., Battaglia, D., Spiegler, A., Deco, G., & Jirsa, V. K. (2015). Functional connectivity dynamics: modeling the switching behavior of the resting state. Neuroimage, 105 , 525 -- 535. Hellyer, P. J., Leech, R., Ham, T. E., Bonnelle, V., & Sharp, D. J. (2013). Individual prediction of white matter injury following traumatic brain injury. Ann Neurol , 73 , 489 -- 499. Hermundstad, A. M., Bassett, D. S., Brown, K. S., Aminoff, E. M., Clewett, D., Freeman, S., Frithsen, A., Johnson, A., Tipper, C. M., Miller, M. B. et al. (2013a). Structural foundations of resting-state and task-based functional connectivity in the human brain. Proceedings of the National Academy of Sciences, 110 , 6169 -- 6174. Hermundstad, A. M., Bassett, D. S., Brown, K. S., Aminoff, E. M., Clewett, D., Freeman, S., Frithsen, A., Johnson, A., Tipper, C. M., Miller, M. B. et al. (2013b). Structural foundations of resting-state and task-based functional connectivity in the human brain. Proceedings of the National Academy of Sciences, 110 , 6169 -- 6174. Hermundstad, A. M., Brown, K. S., Bassett, D. S., Aminoff, E. M., Frithsen, A., Johnson, A., Tipper, C. M., Miller, M. B., Grafton, S. T., & Carlson, J. M. (2014a). Structurally-constrained relationships between cognitive states in the human brain. PLoS Comput Biol , 10 , e1003591. Hermundstad, A. M., Brown, K. S., Bassett, D. S., Aminoff, E. M., Frithsen, A., Johnson, A., Tipper, C. M., Miller, M. B., Grafton, S. T., & Carlson, J. M. (2014b). Structurally-constrained relationships between cognitive states in the human brain. PLoS Comput Biol , 10 , e1003591. Hermundstad, A. M., Brown, K. S., Bassett, D. S., & Carlson, J. M. (2011). Learning, memory, and the role of neural network architecture. PLoS Comput Biol , 7 , e1002063. 16 Honey, C., Sporns, O., Cammoun, L., Gigandet, X., Thiran, J.-P., Meuli, R., & Hagmann, P. (2009). Predicting human resting-state functional connectivity from structural connectivity. Proceedings of the National Academy of Sciences, 106 , 2035 -- 2040. Honey, C. J., Kotter, R., Breakspear, M., & Sporns, O. (2007). Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences, 104 , 10240 -- 10245. Jbabdi, S., Sotiropoulos, S. N., & Behrens, T. E. (2013). The topographic connectome. Current opinion in neurobiology, 23 , 207 -- 215. Johnson, V. E., Stewart, W., & Smith, D. H. (2013). Axonal pathology in traumatic brain injury. Exp Neurol , 246 , 35 -- 43. Johnston, J. M., Vaishnavi, S. N., Smyth, M. D., Zhang, D., He, B. J., Zempel, J. M., Shimony, J. S., Snyder, A. Z., & Raichle, M. E. (2008). Loss of resting interhemispheric functional connectivity after complete section of the corpus callosum. The Journal of neuroscience, 28 , 6453 -- 6458. Kalpinski, R. J., Williamson, M. L., Elliott, T. R., Berry, J. W., Underhill, A. T., & Fine, P. R. (2013). Modeling the prospective relationships of impairment, injury severity, and participation to quality of life following traumatic brain injury. Biomed Res Int, 2013 , 102570. Kandel, E. R., Schwartz, J. H., Jessell, T. M. et al. (2000). Principles of neural science volume 4. McGraw-hill New York. Khambhati, A., Davis, K., Lucas, T., Litt, B., & Bassett, D. S. (2016). Virtual cortical resection reveals push-pull network control preceding seizure evolution. Submitted , . Kinnunen, K. M., Greenwood, R., Powell, J. H., Leech, R., Hawkins, P. C., Bonnelle, V., Patel, M. C., Counsell, S. J., & Sharp, D. J. (2011). White matter damage and cognitive impairment after traumatic brain injury. Brain, 134 , 449 -- 463. Klimm, F., Bassett, D. S., Carlson, J. M., & Mucha, P. J. (2014). Resolving structural variability in network models and the brain. PLOS Comput Biol , 10 , e1003491. Laughlin, S. B. (2001). Efficiency and complexity in neural coding. Novartis Found Symp, 239 , 177 -- 187. Laughlin, S. B., de Ruyter van Steveninck, R. R., & Anderson, J. C. (1998). The metabolic cost of neural information. Nat Neurosci , 1 , 36 -- 41. Lee, S., Ueno, M., & Yamashita, T. (2011). Axonal remodeling for motor recovery after traumatic brain injury requires downregulation of gamma-aminobutyric acid signaling. Cell Death Dis, 2 , e133. Lett, T. A., Voineskos, A. N., Kennedy, J. L., Levine, B., & Daskalakis, Z. J. (2014). Treating working memory deficits in schizophrenia: a review of the neurobiology. Biol Psychiatry, 75 , 361 -- 370. Levy, N., Horn, D., Meilijson, I., & Ruppin, E. (2001). Distributed synchrony in a cell assembly of spiking neurons. Neural Netw , 14 , 815 -- 824. Liu, Y.-Y., Slotine, J.-J., & Barab´asi, A.-L. (2011). Controllability of complex networks. Nature, 473 , 167 -- 173. Luenberger, D. (1979). Introduction to dynamic systems: theory, models, and applications, . Medaglia, J. D., Lynall, M. E., & Bassett, D. S. (2015). Cognitive network neuroscience. J Cogn Neurosci , 27 , 1471 -- 1491. Meunier, D., Achard, S., Morcom, A., & Bullmore, E. (2009). Age-related changes in modular organization of human brain functional networks. Neuroimage, 44 , 715 -- 723. Misi´c, B., Betzel, R. F., Nematzadeh, A., Goni, J., Griffa, A., Hagmann, P., Flammini, A., Ahn, Y.-Y., & Sporns, O. (2015). Cooperative and competitive spreading dynamics on the human connectome. Neuron, 86 , 1518 -- 1529. Muldoon, S. F., Bridgeford, E. W., & Bassett, D. S. (2016a). Small-world propensity and weighted brain networks. Scientific reports, 6 . Muldoon, S. F., Pasqualetti, F., Gu, S., Cieslak, M., Grafton, S. T., Vettel, J. M., & Bassett, D. S. (2016b). Stimulation-based control of dynamic brain networks. arXiv preprint arXiv:1601.00987 , . Muller, F.-J., & Schuppert, A. (2011). Few inputs can reprogram biological networks. Nature, 478 , E4 -- E4. Nenadovic, V., Hutchison, J. S., Dominguez, L. G., Otsubo, H., Gray, M. P., Sharma, R., Belkas, J., & Perez Velazquez, J. L. (2008). Fluctuations in cortical synchronization in pediatric traumatic brain injury. J Neurotrauma, 25 , 615 -- 627. 17 Niven, J. E., & Laughlin, S. B. (2008). Energy limitation as a selective pressure on the evolution of sensory systems. J Exp Biol , 211 , 1792 -- 1804. Nudo, R. J. (2006). Mechanisms for recovery of motor function following cortical damage. Curr Opin Neurobiol , 16 , 638 -- 644. O'Reilly, J. X., Croxson, P. L., Jbabdi, S., Sallet, J., Noonan, M. P., Mars, R. B., Browning, P. G., Wilson, C. R., Mitchell, A. S., Miller, K. L. et al. (2013). Causal effect of disconnection lesions on interhemispheric functional connectivity in rhesus monkeys. Proceedings of the National Academy of Sciences, 110 , 13982 -- 13987. Pasqualetti, F., Zampieri, S., & Bullo, F. (2014). Controllability metrics, limitations and algorithms for complex networks. Control of Network Systems, IEEE Transactions on, 1 , 40 -- 52. Pestilli, F., Yeatman, J. D., Rokem, A., Kay, K. N., & Wandell, B. A. (2014). Evaluation and statistical inference for human connectomes. Nature methods, 11 , 1058 -- 1063. Posner, M. I., & Petersen, S. E. (1989). The attention system of the human brain. Technical Report DTIC Document. Power, J. D., Cohen, A. L., Nelson, S. M., Wig, G. S., Barnes, K. A., Church, J. A., Vogel, A. C., Laumann, T. O., Miezin, F. M., Schlaggar, B. L., & Petersen, S. E. (2011). Functional network organization of the human brain. Neuron, 72 , 665 -- 678. Radhu, N., Daskalakis, Z. J., Guglietti, C. L., Farzan, F., Barr, M. S., Arpin-Cribbie, C. A., Fitzgerald, P. B., & Ritvo, P. (2012). Cognitive behavioral therapy-related increases in cortical inhibition in problematic perfectionists. Brain Stimul , 5 , 44 -- 54. Raichle, M. E. (2015). The brain's default mode network. Annu Rev Neurosci , 38 , 433 -- 447. Raichle, M. E., MacLeod, A. M., Snyder, A. Z., Powers, W. J., Gusnard, D. A., & Shulman, G. L. (2001). A default mode of brain function. Proc Natl Acad Sci U S A, 98 , 676 -- 682. Raichle, M. E., & Snyder, A. Z. (2007). A default mode of brain function: a brief history of an evolving idea. Neuroimage, 37 , 1083 -- 1090. Raja Beharelle, A., Kovacevic, N., McIntosh, A. R., & Levine, B. (2012). Brain signal variability relates to stability of behavior after recovery from diffuse brain injury. Neuroimage, 60 , 1528 -- 1537. Rajan, K., Harvey, C. D., & Tank, D. W. (2016). Recurrent network models of sequence generation and memory. Neuron, 90 , 128 -- 142. Reveley, C., Seth, A. K., Pierpaoli, C., Silva, A. C., Yu, D., Saunders, R. C., Leopold, D. A., & Frank, Q. Y. (2015). Superficial white matter fiber systems impede detection of long-range cortical connections in diffusion mr tractography. Proceedings of the National Academy of Sciences, 112 , E2820 -- E2828. Rusconi, E., Pinel, P., Eger, E., LeBihan, D., Thirion, B., Dehaene, S., & Kleinschmidt, A. (2009). A disconnection account of Gerstmann syndrome: functional neuroanatomy evidence. Ann Neurol , 66 , 654 -- 662. Rushworth, M. F., Behrens, T. E., & Johansen-Berg, H. (2006). Connection patterns distinguish 3 regions of human parietal cortex. Cereb Cortex , 16 , 1418 -- 1430. Salvador, R., Suckling, J., Schwarzbauer, C., & Bullmore, E. (2005). Undirected graphs of frequency-dependent functional connectivity in whole brain networks. Philos Trans R Soc Lond B Biol Sci , 360 , 937 -- 946. Sharp, D. J., Scott, G., & Leech, R. (2014). Network dysfunction after traumatic brain injury. Nat Rev Neurol , 10 , 156 -- 166. Shenoy, K. V., Kaufman, M. T., Sahani, M., & Churchland, M. M. (2011). A dynamical systems view of motor preparation: implications for neural prosthetic system design. Prog Brain Res, 192 , 33 -- 58. Sidaros, A., Engberg, A. W., Sidaros, K., Liptrot, M. G., Herning, M., Petersen, P., Paulson, O. B., Jernigan, T. L., & Rostrup, E. (2008). Diffusion tensor imaging during recovery from severe traumatic brain injury and relation to clinical outcome: a longitudinal study. Sizemore, A., Giusti, C., & Bassett, D. (2015). Classification of weighted networks through mesoscale homological features. arXiv preprint arXiv:1512.06457 , . Sommer, I. E., Slotema, C. W., Daskalakis, Z. J., Derks, E. M., Blom, J. D., & van der Gaag, M. (2012). The treatment of hallucinations in schizophrenia spectrum disorders. Schizophr Bull , 38 , 704 -- 714. 18 Sridharan, D., Levitin, D. J., & Menon, V. (2008). A critical role for the right fronto-insular cortex in switching between central-executive and default-mode networks. Proc Natl Acad Sci U S A, 105 , 12569 -- 125674. Stam, C. J. (2014). Modern network science of neurological disorders. Nat Rev Neurosci , 15 , 683 -- 695. Stephan, K. E., Friston, K. J., & Frith, C. D. (2009). Dysconnection in schizophrenia: from abnormal synaptic plasticity to failures of self-monitoring. Schizophr Bull , 35 , 509 -- 527. Szameitat, A. J., Schubert, T., & Muller, H. J. (2011). How to test for dual-task-specific effects in brain imaging studiesan evaluation of potential analysis methods. Neuroimage, 54 , 1765 -- 1773. Thomas, C., Frank, Q. Y., Irfanoglu, M. O., Modi, P., Saleem, K. S., Leopold, D. A., & Pierpaoli, C. (2014). Anatomical accuracy of brain connections derived from diffusion mri tractography is inherently limited. Proceedings of the National Academy of Sciences, 111 , 16574 -- 16579. Tortella, G., Selingardi, P. M., Moreno, M. L., Veronezi, B. P., & Brunoni, A. R. (2014). Does non-invasive brain stimulation improve cognition in major depressive disorder? a systematic review. CNS Neurol Disord Drug Targets, 13 , 1759 -- 1769. Weiss, S. A., Bassett, D. S., Rubinstein, D., Holroyd, T., Apud, J., Dickinson, D., & Coppola, R. (2011). Functional brain network characterization and adaptivity during task practice in healthy volunteers and people with schizophrenia. Frontiers in human neuroscience, 5 , 81. Yan, G., Ren, J., Lai, Y.-C., Lai, C.-H., & Li, B. (2012). Controlling complex networks: How much energy is needed? Physical review letters, 108 , 218703. Yeo, B. T., Krienen, F. M., Sepulcre, J., Sabuncu, M. R., Lashkari, D., Hollinshead, M., Roffman, J. L., Smoller, J. W., Zollei, L., Polimeni, J. R., Fischl, B., Liu, H., & Buckner, R. L. (2011). The organization of the human cerebral cortex estimated by intrinsic functional connectivity. J Neurophysiol , 106 , 1125 -- 1165. 19
1706.10297
1
1706
2017-06-30T17:59:14
Exploring the Human Connectome Topology in Group Studies
[ "q-bio.NC", "cs.HC" ]
Visually comparing brain networks, or connectomes, is an essential task in the field of neuroscience. Especially relevant to the field of clinical neuroscience, group studies that examine differences between populations or changes over time within a population enable neuroscientists to reason about effective diagnoses and treatments for a range of neuropsychiatric disorders. In this paper, we specifically explore how visual analytics tools can be used to facilitate various clinical neuroscience tasks, in which observation and analysis of meaningful patterns in the connectome can support patient diagnosis and treatment. We conduct a survey of visualization tasks that enable clinical neuroscience activities, and further explore how existing connectome visualization tools support or fail to support these tasks. Based on our investigation of these tasks, we introduce a novel visualization tool, NeuroCave, to support group studies analyses. We discuss how our design decisions (the use of immersive visualization, the use of hierarchical clustering and dimensionality reduction techniques, and the choice of visual encodings) are motivated by these tasks. We evaluate NeuroCave through two use cases that illustrate the utility of interactive connectome visualization in clinical neuroscience contexts. In the first use case, we study sex differences using functional connectomes and discover hidden connectome patterns associated with well-known cognitive differences in spatial and verbal abilities. In the second use case, we show how the utility of visualizing the brain in different topological space coupled with clustering information can reveal the brain's intrinsic structure.
q-bio.NC
q-bio
Exploring the Human Connectome Topology in Group Studies Johnson J.G. Keiriz, Liang Zhan, Morris Chukhman, Olu Ajilore, Alex D. Leow, and Angus G. Forbes 7 1 0 2 n u J 0 3 ] . C N o i b - q [ 1 v 7 9 2 0 1 . 6 0 7 1 : v i X r a Fig. 1. A screen capture of the NeuroCave visualization tool showing the average resting state functional connectome of subjects between the age of 20 and 30 years old for females in the clustering space (left) versus males in the anatomical space (right). Our tool facilitates the simultaneous exploration of multiple connectome datasets in a variety of configurations, enabling researchers to make meaningful comparisons between them and to reason about their differences. Abstract-Visually comparing brain networks, or connectomes, is an essential task in the field of neuroscience. Especially relevant to the field of clinical neuroscience, group studies that examine differences between populations or changes over time within a population enable neuroscientists to reason about effective diagnoses and treatments for a range of neuropsychiatric disorders. In this paper, we specifically explore how visual analytics tools can be used to facilitate various clinical neuroscience tasks, in which observation and analysis of meaningful patterns in the connectome can support patient diagnosis and treatment. We conduct a survey of visualization tasks that enable clinical neuroscience activities, and further explore how existing connectome visualization tools support or fail to support these tasks. Based on our investigation of these tasks, we introduce a novel visualization tool, NeuroCave, to support group studies analyses. We discuss how our design decisions (the use of immersive visualization, the use of hierarchical clustering and dimensionality reduction techniques, and the choice of visual encodings) are motivated by these tasks. We evaluate NeuroCave through two use cases that illustrate the utility of interactive connectome visualization in clinical neuroscience contexts. In the first use case, we study sex differences using functional connectomes and discover hidden connectome patterns associated with well-known cognitive differences in spatial and verbal abilities. In the second use case, we show how the utility of visualizing the brain in different topological space coupled with clustering information can reveal the brain's intrinsic structure. Index Terms-Brain networks, visual comparison, intrinsic topology, connectome visualization 1 INTRODUCTION Over the past decade, the study of the human brain has progressed through advancements in the magnetic resonance imaging (MRI) and • Johnson J.G. Keiriz, Morris Chukhman, Angus G. Forbes, Olu Ajilore, and Alex D. Leow are with University of Illinois at Chicago. E-mail: {jgadel2, chukhman, aforbes}@uic.edu; {oajilore, aleow}@psych.uic.edu • Liang Zhan is with University of Wisconsin–Stout. Email: [email protected] Manuscript received xx xxx. 201x; accepted xx xxx. 201x. Date of Publication xx xxx. 201x; date of current version xx xxx. 201x. For information on obtaining reprints of this article, please send e-mail to: [email protected]. Digital Object Identifier: xx.xxxx/TVCG.201x.xxxxxxx other neuroimaging technology. Those advancements have allowed neu- roscientists to non-invasively probe the brain's structural and functional inter-regional connectivity and derive the human brain connectome [54]. High resolution MRI scans with submillimeter voxel size coupled with advanced non-linear registration algorithms allows the creation of brain label maps [36]. Those maps are created by registering brain MRI scans with a pre-segmented atlas by a highly experienced neuroscientist. Dif- fusion Weighted Imaging (DWI), also called diffusion MRI (dMRI), allows us to reconstruct the brain's white matter fiber tracts through a post-processing procedure called tractography. Counting the fibers interconnecting each pair of regions derived from the label map gener- ates a brain structural connectome. Functional MRI (fMRI) measures the blood-oxygen-level dependent (BOLD) signal which represents the activation level of the different brain regions due to the execution of Fig. 2. Different methods of visualization in connectomics. Top row, left to right: connectome: adjacency matrix (self prepared), node-link [62], edge-bundling [8], connectogram [32]. Bottom row, left to right: tractography: heatmap [41], linked views [33], abstraction [21]. specific tasks. The functional connectome is then generated by com- puting the correlation coefficient of each pair of BOLD signal at the different brain regions. Connectomes are modeled as graphs which allows researchers to use graph theory mathematics to analyze them. Graph metrics can provide insights about the network topological properties such as: its functional integration, which is the network ability to combine infor- mation from its various parts; the clustering or segregation properties, which quantify the existence of groups (clusters or modules); and the network small-world properties, which describe the balance between the functional integration and local clustering. A more extensive review of different graph metrics used in the field of connectomics can be found in [48]. Several important characteristics were derived for the healthy brain network such as small-worldness [1, 49], clustering and modularity [42], and rich-club configuration [58]. The functional con- nectome enables neuroscientists to determine the default mode network (DMN), or default state network, and was found to be the active brain interacting regions when the subject is at wakeful rest and not involved in a task [10]. Group studies are used to study the effects of the onset of neurolog- ical and neuropsychological diseases, as well as age or injury on the human brain, and how it alters the connectome. In a group study, con- nectomes are generated for both healthy and disease populations. Graph metrics are then computed for the resultant networks, and then statisti- cally significant variations in the metrics values can be determined. For a specific disease, researchers will often form a hypothesis based on previous findings in the literature and then attempt to correlate it to the alterations found in the collected graph data metrics [5]. Although the use of graph metrics can provide a detailed aspect of the brain network, it does not deliver potentially useful spatial information, which could help the researcher to contextualize and interpret the results of a graph analysis. This mandates an efficient visualization for the connectome under study in order to help in the interpretation and diagnosis [45]. Moreover, visualization that supports effective comparison becomes crucial in group studies in which alterations in the disease group are more easily understood when analyzed in relation to healthy subjects. The contributions in this paper are as follow: • We delineate visual analytics tasks for clinical neuroscientists, espe- cially as related to group studies. • We introduce a taxonomy of these tasks, providing a comprehensive survey of existing visualization software that supports connectome analysis. • We present NeuroCave, a novel, web-based, immersive visual ana- lytics system that facilitates the visual inspection of structural and functional connectome datasets. With NeuroCave, brain researchers can interact with the connectome in any coordinate system or topo- logical space, as well as group brain regions into different modules on demand. A default side-by-side layout enables simultaneous, syn- chronized manipulation in 3D that facilitates comparison tasks across different subjects or diagnostic groups, or longitudinally within the same subject. • We provide the opportunity for researchers to engage in immersive connectome analytics sessions while wearing portable VR headsets (e.g., Oculus Rift) or on stereoscopic displays (e.g., CAVE systems or 3D video walls). • We reduce visual clutter inherent in dense networks (both in 2D and in 3D views), by introducing a real-time, hardware accelerated edge bundling algorithm, and combining it with a user need-based edge selection strategy. • We present two real-world use cases that demonstrate how our visu- alization system can be used to identify patterns and support compar- ison tasks in order to understand differences between connectome datasets. 2 RELATED WORK The term connectomics was first coined by Sporns et al. [54] to de- scribe the wiring diagram of the anatomical connectivity of the human brain. The interconnectivity found between parcellated brain regions as described above is considered the macroscale connectome. Meso- and micro- scale connectomes are at the local neuronal circuits and single neuronal cells levels respectively, and they currently require an invasive high resolution scans of dissected brains using microtome machinery. Friston defined functional connectivity as the temporal correlation be- tween spatially remote neurophysiological events [22]. This paper focuses on visualization tasks relevant to macroscale functional and structural connectomes, derived from functional and diffusion-weighted MRI respectively. 2.1 Visualization in connectomics Initial efforts into visualizing connectomics data have focused on trac- tography data. The outcome of tractography computation is a set of lines (fibers) spanning the brain white matter area representing the axonal tracts of neuron cells. Volume rendering is used to overlay streamlines on top of the anatomical brain image [17, 44, 51, 52, 66]. Often, a color code is used to identify the direction of the plotted fibers (see Fig. 2). The main task of such visualizations is to allow the user to select and explore specific fiber bundles such as the Corpus Callo- sum interconnecting the left and right hemispheres. Due to the huge number of generated fibers, on the order of 105, a range of techniques are used to facilitate the navigation of the different fiber bundles. Some techniques make use of 2D representations of predefined fiber tracts that are synchronized together with a more conventional 3D visualiza- tion, such as 2D hierarchical tree-like grapha [33], low-dimensional 2D embedding representations [34], and 2D coronal, sagittal and axial pro- jections of the fiber bundles [15] (see Fig. 2). Three dimensional visual abstraction of the fiber bundles is used in the recent work of Everts et al. [21] (see Fig. 2). Heatmaps are also used in some techniques, where they are, for example, overlaid on top of anatomical isosurfaces [41] (see Fig. 2). Segmented neurons and axons in mesoscale connectomes are also visualized using streamlines [2, 6, 25]. Two dimensional adjacency matrices can provide a good overview representation for large connectome datasets (see Fig. 2). Recent work enhances the readability and flexibility of the adjacency matrix [3, 38], but have not been incorporated into software platforms that are readily available for neuroscientists. Moreover, the use of adjacency matrices can hinder users in performing some visual analysis tasks, such as detecting graph alterations in group studies [23, 35]. The most widely used representation for connectome visualization are 3D node-link diagrams, in which nodes are positioned relative to their corresponding anatomical locations, and in which links represent the structural or functional connectivity between the nodes. Many examples of visualization applications are presented in Table 1. An inherent problem of node-link diagrams is the difficulty of effectively displaying the total available links in dense networks. In dense networks, visual clutter due to a preponderance of edge crossings can adversely affect the accuracy and completion time for trend tasks (assessing change in edge weight of a node's connections), connectivity tasks (assessing the connectivity of common neighbors) and region identification tasks (identifying the region with the most changes) [3]. One technique to reduce such clutter is to combine visually compatible edges into bundles according to a compatibility metric, hence the name edge bundling, first introduced by Holten in [28]. Recently, some connectome visualization projects have utilized edge bundling. Two dimensional edge bundling is used by McGraw [40] and Yang et al. [63], while 3D edge bundling is used in [8,9] for representing functional connectivity that shows high levels of common interconnections (see Fig. 2). NeuroCave also utilizes edge bundling in order to reduce visual clutter in large connectome datasets that may have 2500 or more interconnected nodes (see Fig. 8). The connectogram [32], a recent connectome visualization technique, is a node-link diagram in a circular layout (see Fig. 8). Names of brain regions are positioned along the perimeter of a circle, which is split into two halves according to their hemispheric affiliation. Each hemisphere is further divided into the different brain lobes. The inner sphere contains multiple colored rings, each representing a specific metric. The regions are interconnected within the circle using curved lines. This technique prevents some of the clutter that is found in other network visualizations. However, it is harder to correlate anatomical structures with connectivity using a circular diagram like the connectogram, as the multiple rings may make it difficult for a user to make sense of the data [12]. Table 1 contains a list of different visualization tools in the field of connectomics and their capabilities. Although previous efforts have surveyed visualization in connectomics, such as Margulies et al. [39] and Pfister et al. [45], our survey is the first to comprehensively explore recent connectomics visualization software in terms of their ability to support group studies analysis tasks. Moreover, we catalog these software tools by the primary connectome dataset types they support. 2.2 Connectome Analysis Tasks Extending the user task definitions presented in Alper et al. [3] through a thorough investigation of connectome visualization projects and sur- veying neuroscientists that work with group studies, we have identified visual analytics tasks for clinical neuroscientists: T1 Identify regions responsible for specific cognitive functions and study their interactions with other regions. T2 Compare individual networks to the mean or group average con- nectome. In group studies, individual variations as well as joint network characteristics are studied in order to identify commonali- ties or differences. T3 Identify the effect of structural connectivity on the functional activity of the brain. Comparing both structural and functional at the same time to reveal the complex mappings between them [11, 30, 31]. T4 Identify individual or group changes occurring on the structural or functional connectivities due to the onset of disease [50] or aging [20] as well as gender differences. Moreover, researchers can assess its restoration in drug studies [55]. T5 Identify dynamic changes in structural and functional connectivity over time for both within subject and between subjects [16]. T6 Identify structural re-routing occurring after a brain injury or dam- age with rehabilitation training and its effect on functional connec- tivity. Similarly, in case of neurosurgery, it is important to identify affected structural pathways and predict the corresponding loss in motor and cognitive abilities after the procedure. Although each of the the visualization software tools listed in Table 1 may partially address many of these tasks, none provides a visualization that can directly facilitate T2–T5, involving various types of compari- son between datasets, since they all lack the ability to simultaneously load and synchronize a comparative visualization of multiple connec- tomes. The user needs to open multiple instances of the application and usually requires multiple monitors in order to visually be able to compare the structural and functional connectomes of the same subject or two subjects belonging to different groups. Clearly, all user actions will not be synchronized which makes it even harder for assessing visual differences. Applications implemented in scripting languages such as R and MATLAB provide the user with the flexibility to customize views. However, this requires additional efforts as well as programming expertise. Introducing a novel side-by-side layout, NeuroCave helps neuroscientists and researchers efficiently execute T1–T5, which involve comparative analyses, and to simultaneously spot changes occurring within and across subjects. It is important to state that NeuroCave does not target tractography-related usages, such as the sixth task (T6), which, although an important area of connec- tomics visualization, is usually a requirement by neurosurgeons rather than clinical neuroscientists (who are the intended audience for our visualization system). 2.3 Web-based VR While most commonly used visualization tools are dedicated desktop applications, web-based implementations, such as Slice:Drop [24] or BrainBrowser [53], free the user from being attached to a specific operating system [46]. NeuroCave is also a web-based application and runs in any modern browser. The use of stereoscopic techniques can provide a more immersive way to explore brain imaging data [47], Hanel et al. find that healthcare professionals perceive the increased dimensionality provided by stereoscopy as beneficial for understanding depth in the displayed scenery [26]. Moreover, Ware and Mitchell find that the use of stereographic visualizations reduces the perception error rate in graph perception for large graphs with more than 500 nodes [60], and Alper et al. [4] observed that, when coupled with highlighting, stereoscopic representations of 3D graphs outperformed their non- immersive counterpart. NeuroCave lets the user move between desktop and VR environments for interactively exploring 3D connectomes. 2.4 Connectome Topology The techniques of linear and nonlinear dimensionality reduction have been widely used in the field of brain imaging, specially for processing Table 1. Neuroimaging connectomic software. This table indicates the visual analysis tasks (Section 2.2) supported by each of these software tools. Software (V)olume / (S)urface / (G)raph Supported Tasks Description Website Structural DTI Studio S T1, T6 Generates and visualizes tractography and allows user-defined ROI. https://www.mristudio.org/ TrackVis V T1, T6 Generates and visualizes tractography and allows user-defined ROI. http://www.trackvis.org/ Camino V/S T1, T6 Generates and visualizes tractography and allows user-defined ROI. http://camino.cs.ucl.ac.uk/ DoDTI V T1, T6 MATLAB-based tool to generate and visualize tractography. http://neuroimage.yonsei.ac.kr/dodti/ ExploreDTI V/S T1, T6 Generates tractography and allows user-defined ROI. http://www.exploredti.com/ BrainVISA V/S T1. T6 Generates and visualizes tractography data with Python interface. http://brainvisa.info/web/index.html OpenWalnut V/S T1, T6 Visualization of tractography data. http://www.openwalnut.org/ FiberNavigator2 V/S T1, T6 Visualization of tractography data. https://code.google.com/p/fibernavigator2/ TractoR V T1, T6 Plugin to the R programming language to perform tractography and visualize results. http://www.tractor-mri.org.uk/ DSI Studio V/S/G T1, T2, T3, T4, T6 Generates and visualizes tractography and allows user-defined ROI. Generates and visualizes connectomes. http://dsi-studio.labsolver.org/ Brainnetome V/S/G T1, T6 Generates and visualizes tractography and allows user-defined ROI. Generates connectomes. http://diffusion.brainnetome.org/en/latest/ NPerspective S T1 Google-maps like 2D representation of the neural pathways in three projections: Sagittal, Coronal and Axial. http://graphics.cs.brown.edu/research/sciviz/newbraininteraction/ 3D Slicer V/S T1, T6 Generate and visualize tractography data. www.slicer.org Connectome Explorer V/G T1 Visualizes mesoscale connectome generated using MRI scans of dissected brain slices. http://people.seas.harvard.edu/~jbeyer/connectome_explorer.html Functional InstaCorr (SUMA/AFNI) V/S T1, T2, T3, T4 Analysis and visualization of fMRI data. http://afni.nimh.nih.gov/afni VidView S/G T1, T2, T3, T4 Viewer for fMRI connectomic data (connexel). https://github.com/NeuroanatomyAndConnectivity/vidview Brainbundler S/G T1, T2, T3, T4 Visualizes functional connectivity as glyphs on the brain surface. Applies edge bundling to connectome data. https://github.com/NeuroanatomyAndConnectivity/brainbundler/ Fubraconnec S/G T1, T2, T3, T4 Visualization of functional connectome using synchronized anatomy, node-link and circular views. https://code.google.com/p/fubraconnex/ VAMCA V/S T1, T2, T3, T4 MATLAB-based toolbox for analysis and visualization. http://www.nitrc.org/projects/vamca REST G T1, T2, T3, T4 MATLAB-based resting-state functional connectivity analysis and visualization tool. http://restfmri.net/forum/REST CONN V/S/G T1, T2, T3, T4 MATLAB-based for fMRI connectome analysis and visualization. https://www.nitrc.org/projects/conn/ Multimodal Connectome Viewer V/S/G T1, T2, T3, T4 Anatomic and functional connectivity graph visualization. Analysis enabled through Python interface. http://cmtk.org/viewer/ Connectome Workbench V/S/G T1, T2, T3, T4 Investigates data acquired through Human Connectome Project. http://www.humanconnectome.org/software/ BrainNet Viewer V/S/G T1 MATLAB-based brain surface and graph visualization. http://www.nitrc.org/projects/bnv/ Brain Connectivity Toolbox G T2, T3, T4, T5 MATLAB-based list of graph theory analysis functions with basic matrix visualization. https://sites.google.com/site/bctnet/visualization Visual Connectome S/G T1, T2, T3, T4 MATLAB-based structural and functional connectome analysis and visualization. http://code.google.com/p/visualconnectome/ MNET G T1, T2, T3, T4 MATLAB-based dMRI and fMRI processing tool that generates and visualizes tractography data as well as structural and functional connectomes. http://neuroimage.yonsei.ac.kr/mnet/ BRAVIZ V/S T1, T2, T3, T4, T6 MATLAB-based. Overlays dMRI metrics on top of tractography data. http://diego0020.github.io/braviz/ Online BrainBrowser V/S T1, T2 Visualizes structural connectomic data. https://brainbrowser.cbrain.mcgill.ca/ XTK V/S NA Framework for building Web-based neuroimaging viewers. https://github.com/xtk/X Slicedrop V/S T1 Based on XTK, visualizes tractography data. http://slicedrop.com/ Other Connectogram G T1 Standardized circular schematic of multimodal data. http://circos.ca/tutorials/lessons/recipes/cortical_maps/ geometrical properties of platonic solids. In brief, a platonic solid is a regular, convex polyhedron constructed by congruent regular polygonal faces with the same number of faces meeting at each vertex. Five platonic solids exist: tetrahedron, cube, octahedron, dodecahedron and icosahedron, with four, six, eight, twelve, and twenty faces, respectively. Based on how many clusters are generated, a suitable platonic solid is chosen such that its number of faces is greater than the number of these clusters. The glyphs of each cluster are then equally distributed (according to the sunflower algorithm [59]) over half of the spherical cap covering the corresponding face of of a platonic solid embedded in a sphere (see Fig. 3). This enables the user to "enter" into the geometry (i.e. into the "NeuroCave") via one of the unpopulated face(s), provid- ing a more immersive experience of the data. Node visualization. We utilize two different glyphs (spheres and cubes) to differentiate between left and right hemisphere affiliation. Nodes can be colored according to lobar or modular information. Con- trolling nodal transparency is also possible according to their color scheme. Three modes exist: opaque, semi-transparent or totally trans- parent (invisible). Glyphs size is also user adjustable. Edge Visualization. As mentioned in Section 2, edge clutter can be a problem in the visualization of dense node-link diagrams. Our approach to overcome this problem is based on two steps. First, we provide the option to hide all edges by default (i.e., to show only the nodes). The user can then select a root node, and all connected edges stemming from this node will be displayed. Second, to overcome the clutter occurring from edge crossings, we use the force directed edge bundling (FDEB) algorithm to group edges going in the same direction [29]. FDEB is an iterative algorithm that consists of cycles. In each cycle, we subdivide an edge into a specified number of points (we chose 6 cycles, and we double the number of points each cycle, ending up with 64 subdivision points plus the two original points of the edge). After the subdivision, we iteratively move each subdivision point in an update step to a new position determined by modeling the forces among the points. Each point is affected by the sum of spring and electrostatic forces (Fs and Fe). For an edge i, at a subdivision point pi j, Fs is defined as: Fs = kp((cid:107)pi( j−1) − pi j(cid:107) +(cid:107)pi j − pi( j+1)(cid:107)) (1) where kp is a spring constant and pi( j−1) and pi( j+1) are the neighbor- ing points of pi j, the (cid:107).(cid:107) is the length. The electrostatic force Fe is defined as Fe = ∑ m∈E 1 (cid:107)pi j − pm j(cid:107) (2) where E is a set of compatible edges with the current edge, pm j is the corresponding subdivision point to point pi j in the edges belonging to the set E. The compatibility metrics used to define the set E are defined in details in Holten et al. [29]. Each cycle contains a prespecified number of update steps. Our original Javascript implementation of NeuroCave turned out to be too slow for large numbers of edges (more than 500), preventing a real-time experience. Therefore, we used (and enhanced) a WebGL texture-based implementation suggested by Wu et al. [61]. The FDEB algorithm is parallelizable since the division and update operations are performed on each point independently. The texture-based method stores the subdivision points in a 2D GPU texture, where each row represents the 3D coordinates of points belonging to the same edge. Since write operations are unavailable to GPU textures in WebGL, a ping-pong algorithm 4 is used to write results to a framebuffer object (FBO). Two shaders are utilized: the first performs the subdivision operation, and the second executes the update steps. In Wu et al.'s implementation, the limitation on the texture size limits is not addressed. GPU textures possess a limitation on their sizes. Hence, a large number of edges can not be fit in one texture. We enhanced the algorithm using tiling. Since the total number of points of each edge after all cycles will be known ahead (64 subdivision + 2 end points), we tile the edges when the maximum number of possible rows per texture is achieved. 4https://www.khronos.org/opengl/wiki/Memory_Model. Fig. 3. The figure shows the proposed method of distributing clustered nodes. The nodes are distributed on half the spherical cap covering a platonic solid face. The chosen platonic solid must have a number of faces that is equal to or larger than the number of clusters (for 8 clusters a dodecahedron was chosen). fMRI data [13, 57]. Recently, nonlinear dimensionality reduction tech- niques, such as isomap and t-distributed stochastic neighbor embedding (t-SNE), have been applied onto the human connectome in order to find the brain's intrinsic geometry [64]. Using such embedding, new 3D topologies for the brain were found other than the regular anatomical one. Such topologies depends on the internodal graph shortest path length computed using Dijkstra's algorithm [19]. Nodes possessing efficient connectivity to the rest of the brain were found to be positioned near the origin of those new topologies inferring their importance. Neu- roCave facilitates finding topological alterations in group studies. This helps researchers in studying the connectome topology of healthy sub- jects, such as the DMN, and comparing it with diseased groups. In DMN, important hub nodes, such as the precuneus (see Fig. 8), possess a strong connectivity with the rest of the brain and are located near the center of the brain's intrinsic geometry [64]. A deviation from such topology in disease groups need to be studied for illnesses associated with DMN alterations such as Alzheimer's disease which is known to disrupt it [10]. 3 THE NeuroCave SOFTWARE SYSTEM NeuroCave is implemented as a web-based application. The Javascript WebGL-based graphics library three.js 1 was used for the 3D rendering and visualization functionalities. It runs on all major web browsers, and is thus platform independent. The default view is formed of two side-by-side 3D rendering views (see Fig. 1). Each view allows the interactive visualization of a connectome as a node-link diagram. Group Visualization. The application loads subjects data from a study folder. The folder should contains all adjacency matrices as well as the corresponding topological and clustering information of the sub- jects within the study. An indexing file states the subject ID and its corresponding data files. Each study requires a predefined Atlas that provides numerical labels and their corresponding anatomical names to each node. NeuroCave directly supports three Atlases: FSL-based which consists of 82 labels of FreeSurfer 2, brain hierarchical Atlas (BHA) made of 2514 labels [18] and Harvard-Oxford Atlas made of 177 labels.3 However, additional Atlases can be easily created follow- ing the pre-existing ones as example. Topology Visualization. The application layouts the nodes according to the provided topological information. Topologies can be the anatom- ical positioning or an applied transformation in some abstract space as explained in Section 2. The available topologies are automatically identified by the application. Clustering Visualization. Clustering information is input as a vector of integer values. Each value represents a different module or cluster. Since there are no prespecified positions for clusters, we exploit the 1http://threejs.org. 2https://surfer.nmr.mgh.harvard.edu/. 3http://neuro.imm.dtu.dk/wiki/Harvard-Oxford_Atlas. Fig. 4. Edge bundling (left) versus no edge bundling (right) in clustering space displaying four clusters. Notice the clutter caused by the edges in the right panel versus the edge bundling case. Also, notice the edge color gradient that depends on the nodes color and their relative strengths: the orange sphere possesses a greater strength compared to the blue nodes it is connected to, while it possesses a lesser strength compared to the green nodes it is connected to. NeuroCave enables users to toggle between edge bundling and edge coloring modes on demand. Our texture-based implementation can bundle the closet 1000 edges to the selected node at interactive rates on a desktop computer (Intel Core i7, 3.4 GHz CPU, Nvidia GTX 1070 GPU card and 32GB RAM). Each edge is colored according to a two color gradient chosen ac- cording to the interconnected nodes colors. The gradient is skewed towards the node possessing the higher nodal strength (the sum of weights of links connected to the node) value (see Fig. 4). For an edge made of n points the color, Cout, at point i is Cout = C2 + (C1 −C2)R (3) 2))/P1P2 R = (r2(p2 − 0.5) + r(0.5− p2 (4) where C1 and C2 are the nodes, r = i/(n − 1), P1 = S1/S1 + S2), P2 = 1− P1, S1 and S2 are the nodes strengths. This allows the user to recognize the strength of the selected node with respect to its intercon- nected neighbors which helps identifying strong and important nodes and hubs as well as highlight the reason for modular changes when occurred in group studies as will be shown in Section 4. Virtual Reality. In addition to the standard 3D manipulations of pan- ning, rotating, and zooming, NeuroCave supports the Oculus Rift and the Oculus Rift Touch controllers. The Touch controllers are a pair of VR input devices that track each hand, enabling an effective gesture- based manipulation of the VR environment. The user selects the pre- view area to be explored in VR and then uses the thumbsticks of the Touch devices to navigate the visualized connectome through panning, rotating, and zooming. Nodal selection is enabled via a two step proce- dure: first, pressing the grip button lets the user point at and highlight a node; second, pressing the index button selects the highlighted node. 4 RESULTS AND DISCUSSION In this section, we present two use cases that demonstrate how clinical neuroscientists can perform the tasks described in Section 2. 4.1 Use Case 1 Our first use case investigates the sex-specific resting-state functional connectomes in the F1000 repository, a large 986 subjects publicly- available resting-state fMRI connectome dataset.5 The following post- processing was performed: first, to eliminate the potential confounding effect of age, we only included subjects between 20 to 30 years old (319 females at 23.25 ± 2.26 years of age and 233 males at 23.19 ± 2.35). We then constructed hierarchical modularity in the form of 4-level dendrograms (16 clusters at the most local level) [65]. Previ- ously, rigorous permutation testing established sex differences already statistically significant at the first level (2 clusters; p = 0.0378) for the following six regions: posterior part of left and right frontal pole (near the fronto-temporal junction), left precentral gyrus, right hippocampus, right superior temporal gyrus posterior division (sTG-p) and the right inferior frontal gyrus pars opercularis (POP) (for more detail about the statistical analysis, refer to [65]). In order to interpret these sex differences, male and female group- averaged connectomes were visualized in NeuroCave. We highlighted the connectivity of the aforementioned six regions in both the anatomi- cal and clustering (1st hierarchical level) spaces and applied a threshold to restrict edges to values with absolute correlation values for fMRI BOLD signals larger than 0.1 (see Fig. 5, Fig. 6). From the clustering space, affiliation patterns for the right hippocampus and sTG-p differ in that in women they are clustered with other frontal and temporal ROIs as part of the default mode network (DMN), while in males they are clustered with other parietal and occipital non-DMN ROIs. Notably, the affiliation differences are opposite for the left and right frontal poles, the left precentral gyrus and the right POP, such that they are more clustered along with several non-DMN parietal and occipital ROIs in the average female connectome (see Fig. 5). The left and right fronto-temporal junctions are part of the larger language system, with the right POP (functionally coupled with its homologous area on the left that forms the Broca's language area) linked to the processing of semantic information [27], and the superior temporal gyrus involved in the comprehension of language (as well as in the perception of emotions in facial stimuli [7]). Thus, the observed connectivity differences are related to well-known sex differences in language and emotion/affect processing, as well as differences in self- referential/autobiographical information retrieval. By contrast, the 5https://www.nitrc.org/projects/fcon_1000/. Fig. 5. Left and right frontal poles (red and green rings encircling the sphere and cube glyphs), left precentral gyrus (yellow ring) and the right inferior frontal gyrus (blue ring) nodes selected in anatomical (left panel) and clustering (right panel) spaces for female and male average connectomes. The color code represents the two clusters in the anatomical space, while in the clustering space, it represents the lobe affiliation. Notice the change in group affiliation of the four nodes between the female and male connectomes (left panel). Fig. 6. Right hippocampus (red ring) and superior temporal gyrus (green ring) nodes selected in anatomical (left panel) and clustering (right panel) spaces for female and male average connectomes. Color code is similar to the above figure. Notice the change in group affiliation of the four nodes between the female and male connectomes (left panel). Fig. 7. Right hippocampus (left panel) and left hippocampus (right panel) nodes selected in the anatomical space for female and male average connectomes. Note the tendency in females for both left and right hippocampus to be functionally connected to the contralateral frontal lobe (but not in males). Note that such differences are not previously known to our clinicians and neuroscientists. Here, the color code represents connectome's hierarchical modularity, represented as a dendrogram, at the most global level (2 modules). Fig. 8. Connectivity emerging from the anterior (red ring) and posterior (yellow ring) parts of the precuneus in anatomical (left panel) and isomap "intrinsic" (middle panel) spaces. Right panel: Connectivity emerging from the posterior part of the precuneus visualized in the intrinsic space. The color code represents the modular structure of the connectome consisting of 4 modules. Note that the orange community contains the default mode network. The inset plot shows the residual geodesics for the first 10 dimensions of the isomap dimensionality reduction algorithm. hippocampus is known to play an important role in the formation of new memory, retrieval of declarative long-term memory, and the management and processing of spatial and spatiotemporal working memory. The modular affinity between the right hippocampus and other non-DMN regions in the parietal and occipital lobes in males may thus be related to their established advantage in spatial tasks, including spatial visualization, perception and mental rotation [37] (visual system is heavily composed of the occipital lobe responsible for first-level visual processing, while part of the parietal lobe is instrumental for visuospatial skills). Lastly, closer inspections of the hippocampal functional connec- tivity between women and men further revealed differences that are not known previously to our clinicians (Fig 7). Indeed, in the female connectome there is a strong tendency for both left and right hippocam- pus to functionally communicate with the contralateral frontal lobe, a phenomenon that is not visually present in the male connectome. The discovery of such subtle differences is dependent upon an iteratively explorative visualization process, only possible with the comprehensive suite of tools implemented in NeuroCave. This use case demonstrates the effectiveness of NeuroCave is sup- porting tasks T3 and T4, enabling neuroscientists to better understand neurological gender differences in connectome datasets and to observe how these differences relate to various psychological studies. 4.2 Use Case 2 Our second use case explores a resting-state fMRI high-resolution dataset consisting of 2514 regions-of-interest publicly available at NI- TRC.6 Most functional connectome studies treat the negative corre- lation entries (anticorrelated BOLD signals) in the network by either taking the absolute value or clamping to zero (heuristically chosen) which affects all network metrics' computations. Instead, we applied a recently proposed probabilistic framework which more rigorously ac- counts for negative edges [65]. For an N×N functional connectome, the framework estimates the probability that an edge, ei j, is positive or neg- ative using the connectomes of a group of subjects (i and j varies from 1 to N). The edge positivity EPi j and edge negativity ENi j form a com- plementary pair, since EPi j + ENi j = 1, and thus can be jointly coded using the angle of a unit-length vector: θi j = arctan((cid:112)ENi j/EPi j) which varies from 0 to 90 degrees. Using dissimilarity graph embed- ding, each node, i, is then embedded in an n-dimensional space at the coordinates (θi1,θi2, ...θiN )T . In order to reduce the dimensionality of the resultant topology, we applied the nonlinear dimensionality reduc- tion isomap algorithm [56] that aims to preserve geodesic distances in 6http://www.nitrc.org/frs/?group_id=964. a lower-dimensional space (i.e., the "intrinsic space"). Separately, we determined the community structure by maximizing the Q-modularity metric [43]. From the inset plot in Fig. 8, it is clear that the degree of isometric embedding levels off after the fifth dimension, thus suggesting that the intrinsic topology of the resting-state functional connectome has a di- mension of five, a novel finding that merits further research. To enable 3D visualization of the transformed topology, we retained the first three dimensions of isomap and visualized the modular structure of the brain in both the anatomical space as well as this novel intrinsic topological space (Fig 8). As expected, the nodes assigned to the same community are positioned close to one another in this novel topology. To illustrate how neuroscientists explore this complex topological space and gain further insight into the brain, we selected two nodes that belong to the anterior and posterior part of the precuneus. Although the nodes are anatomically close to each other, they are known to be functionally distinct (thus belong to different modules). Indeed, the anterior part of the precuneus is an important region of the DMN known to be respon- sible for self-referential imagery (thinking about self) and is involved in autobiographical tasks and self-consciousness, thus activated during "resting consciousness" [14]. As we can see in Fig. 8, in this intrinsic space the anterior part of the precuneus, while assigned to the orange module (DMN), exhibits diverse connections with various regions of the brain: the sensori-motor module (blue) and the frontoparietal exec- utive or task-positive system (red). By contrast, the posterior precuneus is part of the visual system (green) and has a relatively restricted pattern (compared with its more anterior counterpart) of connectivity with the rest of the brain. Notably, such connectivity differences only become visually apparent when neuroscientists visualize in this novel space. This use case highlights how our visualization system is effective at supporting tasks T1 and T2, enabling neuroscientists to explore high density connectomics data comprising a large number of ROIs in order to identify and further understand the specific functionality of different brain regions. Moreover, the side-by-side visualization enables users to reason about the relationships between the anatomical and the intrinsic topology, facilitating further insight into how the same brain region can take part in different tasks. 5 CONCLUSION AND FUTURE WORK In this paper, we presented NeuroCave, a novel VR-compatible visual- ization environment for exploring and analyzing the human connectome. We also performed a task taxonomy for neuroscientists and researchers in the field of connectomics. NeuroCave facilitates comparison tasks of two connectomes, an activity that clinical neuroscientists often require for the analysis of group studies. Our system includes various visu- alization enhancements, such as edge-bundling to reduce edge clutter and edge gradient coloring to help identify nodal strengths. More- over, it enables users to explore connectomic data in an immersive VR environment, which can help to improve their perception of the data under study. The GPU is extensively used in NeuroCave for hardware acceleration of both computation and visualization tasks. We demon- strated the effectiveness of the system using two real-world use cases in which neuroscientists were able to use NeuroCave to effectively perform research and analysis tasks. However, many challenges in visualizing the brain connectome remain. Future work will adapt our system to support temporally- varying dynamic connectome datasets (T5). We also will continue to work with clinical neuroscientists to integrate additional analytic tools to assist with diagnosis and treatment of patients. Finally, our side-by- side visualization framework can be extended to other domains that require group study analysis, such as cancer biology, where researchers are interested in comparing experiment vs. control networks or disease vs. normal gene networks in order to better understanding cancer and other disease genomes. REFERENCES [1] S. Achard, R. Salvador, B. Whitcher, J. Suckling, and E. Bullmore. A resilient, low-frequency, small-world human brain functional network with highly connected association cortical hubs. Journal of Neuroscience, 26(1):63–72, 2006. [2] A. K. Ai-Awami, J. Beyer, D. Haehn, N. Kasthuri, J. W. Lichtman, H. Pfis- ter, and M. Hadwiger. NeuroBlocks–Visual tracking of segmentation and proofreading for large connectomics projects. IEEE Transactions on Visualization and Computer Graphics, 22(1):738–746, 2016. [3] B. Alper, B. Bach, N. Henry Riche, T. Isenberg, and J.-D. Fekete. Weighted graph comparison techniques for brain connectivity analysis. In Proceed- ings of the SIGCHI Conference on Human Factors in Computing Systems, pp. 483–492, 2013. [4] B. Alper, T. Hollerer, J. Kuchera-Morin, and A. G. Forbes. Stereoscopic highlighting: 2d graph visualization on stereo displays. IEEE Transactions on Visualization and Computer Graphics, 17(12):2325–2333, 2011. [5] D. A. Angulo, C. Schneider, J. H. Oliver, N. Charpak, and J. T. Hernandez. A multi-facetted visual analytics tool for exploratory analysis of human brain and function datasets. Frontiers in Neuroinformatics, 10, 2016. [6] J. Beyer, A. Al-Awami, N. Kasthuri, J. W. Lichtman, H. Pfister, and M. Hadwiger. ConnectomeExplorer: Query-guided visual analysis of large volumetric neuroscience data. IEEE Transactions on Visualization and Computer Graphics, 19(12):2868–2877, 2013. [7] E. D. Bigler, S. Mortensen, E. S. Neeley, S. Ozonoff, L. Krasny, M. John- son, J. Lu, S. L. Provencal, W. McMahon, and J. E. Lainhart. Superior temporal gyrus, language function, and autism. Developmental Neuropsy- chology, 31(2):217–238, 2007. [8] J. Bottger, A. Schafer, G. Lohmann, A. Villringer, and D. S. Margulies. Three-dimensional mean-shift edge bundling for the visualization of func- tional connectivity in the brain. IEEE Transactions on Visualization and Computer Graphics, 20(3):471–480, 2014. [9] J. Bottger, R. Schurade, E. Jakobsen, A. Schaefer, and D. S. Margulies. Connexel visualization: A software implementation of glyphs and edge- bundling for dense connectivity data using brainGL. Frontiers in Neuro- science, 8:15, 2014. [10] R. L. Buckner, J. R. Andrews-Hanna, and D. L. Schacter. The brain's default network. Annals of the New York Academy of Sciences, 1124(1):1– 38, 2008. [11] E. Bullmore and O. Sporns. Complex brain networks: Graph theoretical analysis of structural and functional systems. Nature Reviews Neuro- science, 10(3):186–198, 2009. [12] M. Burch and D. Weiskopf. On the benefits and drawbacks of radial diagrams. In Handbook of Human-Centric Visualization, pp. 429–451. 2014. [13] V. D. Calhoun, J. Liu, and T. Adalı. A review of group ica for fmri data and ica for joint inference of imaging, genetic, and erp data. Neuroimage, 45(1):S163–S172, 2009. [14] A. E. Cavanna. The precuneus and consciousness. CNS Spectrums, 12(07):545–552, 2007. [15] W. Chen, Z. Ding, S. Zhang, A. MacKay-Brandt, S. Correia, H. Qu, J. A. Crow, D. F. Tate, Z. Yan, and Q. Peng. A novel interface for interactive exploration of DTI fibers. IEEE Transactions on Visualization and Computer Graphics, 15(6):1433–1440, 2009. [16] N. A. Crossley, T. R. Marques, H. Taylor, C. Chaddock, F. DellAcqua, A. A. Reinders, V. Mondelli, M. DiForti, A. Simmons, A. S. David, et al. Connectomic correlates of response to treatment in first-episode psychosis. Brain, p. aww297, 2016. [17] M. Da Silva, S. Zhang, C. Demiralp, and D. H. Laidlaw. Visualizing diffusion tensor volume differences. In IEEE Visualization 01 Proceedings, Work in Progress, 2001. [18] I. Diez, P. Bonifazi, I. n. Escudero, B. Mateos, M. A. Munoz, S. Stramaglia, and J. M. Cortes. A novel brain partition highlights the modular skeleton shared by structure and function. Scientific Reports, 5, 2015. [19] E. W. Dijkstra. A note on two problems in connexion with graphs. Nu- merische Mathematik, 1(1):269–271, 1959. [20] N. U. Dosenbach, B. Nardos, A. L. Cohen, D. A. Fair, J. D. Power, J. A. Church, S. M. Nelson, G. S. Wig, A. C. Vogel, C. N. Lessov- Schlaggar, et al. Prediction of individual brain maturity using fMRI. Science, 329(5997):1358–1361, 2010. [21] M. H. Everts, E. Begue, H. Bekker, J. B. Roerdink, and T. Isenberg. Exploration of the brains white matter structure through visual abstrac- tion and multi-scale local fiber tract contraction. IEEE Transactions on Visualization and Computer Graphics, 21(7):808–821, 2015. [22] K. J. Friston. Functional and effective connectivity in neuroimaging: A synthesis. Human Brain Mapping, 2(1-2):56–78, 1994. [23] M. Ghoniem, J.-D. Fekete, and P. Castagliola. On the readability of graphs using node-link and matrix-based representations: A controlled experiment and statistical analysis. Information Visualization, 4(2):114–135, 2005. [24] D. Haehn. Slice:Drop: Collaborative medical imaging in the browser. In ACM SIGGRAPH 2013 Computer Animation Festival. ACM, 2013. [25] D. Haehn, S. Knowles-Barley, M. Roberts, J. Beyer, N. Kasthuri, J. W. Lichtman, and H. Pfister. Design and evaluation of interactive proofreading tools for connectomics. IEEE transactions on visualization and computer graphics, 20(12):2466–2475, 2014. [26] C. Hanel, P. Pieperhoff, B. Hentschel, K. Amunts, and T. Kuhlen. Interac- tive 3D visualization of structural changes in the brain of a person with corticobasal syndrome. Frontiers in Neuroinformatics, 8(42), 2014. [27] S. Heim, K. Alter, A. K. Ischebeck, K. Amunts, S. B. Eickhoff, H. Mohlberg, K. Zilles, D. Y. von Cramon, and A. D. Friederici. The role of the left Brodmann's areas 44 and 45 in reading words and pseudowords. Cognitive Brain Research, 25(3):982–993, 2005. [28] D. Holten. Hierarchical edge bundles: Visualization of adjacency relations in hierarchical data. IEEE Transactions on Visualization and Computer Graphics, 12(5):741–748, 2006. [29] D. Holten and J. J. Van Wijk. Force-directed edge bundling for graph visualization. In Proceedings of the 11th Eurographics / IEEE - VGTC Conference on Visualization, EuroVis'09, pp. 983–998. Wiley Online Library, 2009. [30] C. Honey, O. Sporns, L. Cammoun, X. Gigandet, J.-P. Thiran, R. Meuli, and P. Hagmann. Predicting human resting-state functional connectivity from structural connectivity. Proceedings of the National Academy of Sciences, 106(6):2035–2040, 2009. [31] C. J. Honey, R. Kotter, M. Breakspear, and O. Sporns. Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences, 104(24):10240–10245, 2007. [32] A. Irimia, M. C. Chambers, C. M. Torgerson, and J. D. Van Horn. Circular representation of human cortical networks for subject and population-level connectomic visualization. Neuroimage, 60(2):1340–1351, 2012. [33] R. Jianu, C. Demiralp, and D. Laidlaw. Exploring 3d dti fiber tracts with linked 2d representations. IEEE Transactions on Visualization and Computer Graphics, 15(6):1449–1456, 2009. [34] R. Jianu, C. Demiralp, and D. H. Laidlaw. Exploring brain connectivity with two-dimensional neural maps. IEEE Transactions on Visualization and Computer Graphics, 18(6):978–987, 2012. [35] R. Keller, C. M. Eckert, and P. J. Clarkson. Matrices or node-link diagrams: which visual representation is better for visualising connectivity models? Information Visualization, 5(1):62–76, 2006. [36] A. Klein, J. Andersson, B. A. Ardekani, J. Ashburner, B. Avants, M.- C. Chiang, G. E. Christensen, D. L. Collins, J. Gee, P. Hellier, et al. Evaluation of 14 nonlinear deformation algorithms applied to human brain MRI registration. Neuroimage, 46(3):786–802, 2009. [37] M. C. Linn and A. C. Petersen. Emergence and characterization of sex differences in spatial ability: A meta-analysis. Child Development, pp. 1479–1498, 1985. approach for interactively visualizing large graphs. In Big Data (Big Data), 2015 IEEE International Conference on, pp. 2501–2508, 2015. [62] M. Xia, J. Wang, and Y. He. Brainnet viewer: a network visualization tool for human brain connectomics. PloS one, 8(7):e68910, 2013. [63] X. Yang, L. Shi, M. Daianu, H. Tong, Q. Liu, and P. Thompson. Blockwise human brain network visual comparison using NodeTrix representation. IEEE Transactions on Visualization and Computer Graphics, 23(1):181– 190, 2017. [64] A. Q. Ye, O. A. Ajilore, G. Conte, J. GadElkarim, G. Thomas-Ramos, L. Zhan, S. Yang, A. Kumar, R. L. Magin, A. G. Forbes, and A. D. Leow. The intrinsic geometry of the human brain connectome. Brain Informatics, 2(4):197–210, 2015. [65] L. Zhan, L. M. Jenkins, O. Wolfson, J. Jonaris, K. N. GadElkarim, P. M. Thompson, O. Ajilore, K. Moo, and A. L. Chung. The importance of being negative: A serious treatment of non-trivial edges in brain functional connectome. arXiv preprint arXiv:1609.01384v2, 2016. [66] S. Zhang, C. Demiralp, and D. H. Laidlaw. Visualizing diffusion tensor mr images using streamtubes and streamsurfaces. IEEE Transactions on Visualization and Computer Graphics, 9(4):454–462, 2003. [38] C. Ma, R. V. Kenyon, A. G. Forbes, T. Berger-Wolf, B. J. Slater, and D. A. Llano. Visualizing dynamic brain networks using an animated dual-representation. In Proceedings of the Eurographics Conference on Visualization (EuroVis), pp. 73–77, 2015. [39] D. S. Margulies, J. Bottger, A. Watanabe, and K. J. Gorgolewski. Visualiz- ing the human connectome. NeuroImage, 80:445–461, 2013. [40] T. McGraw. Graph-based visualization of neuronal connectivity using matrix block partitioning and edge bundling. In International Symposium on Visual Computing, pp. 3–13, 2015. [41] T. McGraw and M. Nadar. Stochastic DT-MRI connectivity mapping on the GPU. IEEE Transactions on Visualization and Computer Graphics, 13(6), 2007. [42] D. Meunier, R. Lambiotte, and E. T. Bullmore. Modular and hierarchically modular organization of brain networks. Frontiers in Neuroscience, 4:200, 2010. [43] M. E. Newman. Modularity and community structure in networks. Pro- ceedings of the National Academy of Sciences, 103(23):8577–8582, 2006. [44] V. Petrovic, J. Fallon, and F. Kuester. Visualizing whole-brain DTI tractog- raphy with GPU-based tuboids and LoD management. IEEE Transactions on Visualization and Computer Graphics, 13(6):1488–1495, 2007. [45] H. Pfister, V. Kaynig, C. P. Botha, S. Bruckner, V. J. Dercksen, H.-C. Hege, and J. B. Roerdink. Visualization in connectomics. In Scientific Visualization, pp. 221–245. 2014. [46] C. Pieloth, J. M. Pizarro, T. Knosche, B. Maess, and M. Fuchs. An online system for neuroelectromagnetic source imaging. In IEEE International Conference on Intelligent Data Acquisition and Advanced Computing Systems (IDAACS), vol. 1, pp. 270–274, 2013. [47] G. M. Rojas, M. G´alvez, N. Vega Potler, R. C. Craddock, D. S. Margulies, F. X. Castellanos, and M. P. Milham. Stereoscopic three-dimensional visualization applied to multimodal brain images: Clinical applications and a functional connectivity atlas. Frontiers in Neuroscience, 8:328, 2014. [48] M. Rubinov and O. Sporns. Complex network measures of brain connec- tivity: Uses and interpretations. Neuroimage, 52(3):1059–1069, 2010. [49] R. Salvador, J. Suckling, M. R. Coleman, J. D. Pickard, D. Menon, and E. Bullmore. Neurophysiological architecture of functional magnetic resonance images of human brain. Cerebral Cortex, 15(9):1332–1342, 2005. [50] E. J. Sanz-Arigita, M. M. Schoonheim, J. S. Damoiseaux, S. A. Rom- bouts, E. Maris, F. Barkhof, P. Scheltens, and C. J. Stam. Loss of small- worldnetworks in alzheimer's disease: graph analysis of fmri resting-state functional connectivity. PloS One, 5(11):e13788, 2010. [51] T. Schultz, H. Theisel, and H.-P. Seidel. Topological visualization of brain diffusion mri data. IEEE Transactions on Visualization and Computer Graphics, 13(6), 2007. [52] A. Sherbondy, D. Akers, R. Mackenzie, R. Dougherty, and B. Wandell. Exploration of the brains white matter pathways with dynamic queries. IEEE Transactions on Visualization and Computer Graphics, 11(4):419– 430, 2005. [53] T. Sherif, N. Kassis, M.- ´E. Rousseau, R. Adalat, and A. C. Evans. Brain- browser: distributed, web-based neurological data visualization. Frontiers in Neuroinformatics, 8:89, 2015. [54] O. Sporns, G. Tononi, and R. Kotter. The human connectome: A structural description of the human brain. PLoS Computational Biology, 1(4):e42, 2005. [55] R. Tadayonnejad, O. Ajilore, B. J. Mickey, N. A. Crane, D. T. Hsu, A. Ku- mar, J.-K. Zubieta, and S. A. Langenecker. Pharmacological modulation of pulvinar resting-state regional oscillations and network dynamics in major depression. Psychiatry Research: Neuroimaging, 252:10–18, 2016. [56] J. B. Tenenbaum, V. De Silva, and J. C. Langford. A global ge- ometric framework for nonlinear dimensionality reduction. Science, 290(5500):2319–2323, 2000. [57] B. Thirion and O. Faugeras. Nonlinear dimension reduction of fMRI data: The Laplacian embedding approach. In Proceedings of the IEEE International Symposium on Biomedical Imaging, pp. 372–375, 2004. [58] M. P. Van Den Heuvel and O. Sporns. Rich-club organization of the human connectome. Journal of Neuroscience, 31(44):15775–15786, 2011. [59] H. Vogel. A better way to construct the sunflower head. Mathematical biosciences, 44(3-4):179–189, 1979. [60] C. Ware and P. Mitchell. Visualizing graphs in three dimensions. ACM Transactions on Applied Perception (TAP), 5(1):2, 2008. [61] J. Wu, L. Yu, and H. Yu. Texture-based edge bundling: A web-based
1812.06234
1
1812
2018-12-15T05:15:40
Integrating Artificial Intelligence with Real-time Intracranial EEG Monitoring to Automate Interictal Identification of Seizure Onset Zones in Focal Epilepsy
[ "q-bio.NC", "cs.AI", "q-bio.QM" ]
An ability to map seizure-generating brain tissue, i.e., the seizure onset zone (SOZ), without recording actual seizures could reduce the duration of invasive EEG monitoring for patients with drug-resistant epilepsy. A widely-adopted practice in the literature is to compare the incidence (events/time) of putative pathological electrophysiological biomarkers associated with epileptic brain tissue with the SOZ determined from spontaneous seizures recorded with intracranial EEG, primarily using a single biomarker. Clinical translation of the previous efforts suffers from their inability to generalize across multiple patients because of (a) the inter-patient variability and (b) the temporal variability in the epileptogenic activity. Here, we report an artificial intelligence-based approach for combining multiple interictal electrophysiological biomarkers and their temporal characteristics as a way of accounting for the above barriers and show that it can reliably identify seizure onset zones in a study cohort of 82 patients who underwent evaluation for drug-resistant epilepsy. Our investigation provides evidence that utilizing the complementary information provided by multiple electrophysiological biomarkers and their temporal characteristics can significantly improve the localization potential compared to previously published single-biomarker incidence-based approaches, resulting in an average area under ROC curve (AUC) value of 0.73 in a cohort of 82 patients. Our results also suggest that recording durations between ninety minutes and two hours are sufficient to localize SOZs with accuracies that may prove clinically relevant. The successful validation of our approach on a large cohort of 82 patients warrants future investigation on the feasibility of utilizing intra-operative EEG monitoring and artificial intelligence to localize epileptogenic brain tissue.
q-bio.NC
q-bio
Integrating Artificial Intelligence with Real-time Intracranial EEG Monitoring to Automate Interictal Identification of Seizure Onset Zones in Focal Epilepsy Authors Yogatheesan Varatharajah1, Brent Berry2,5, Jan Cimbalnik3, Vaclav Kremen2,3, Jamie Van Gompel4, Matt Stead2, Benjamin Brinkmann2,5, Ravishankar Iyer1, and Gregory Worrell2 1 Electrical and Computer Engineering, University of Illinois, Urbana, IL 61801, USA. 2 Mayo Systems Electrophysiology Laboratory, Department of Neurology, Mayo Clinic, Rochester, MN 55905, USA 3 Czech Institute of Informatics, Robotics and Cybernetics, Czech Technical University in Prague, 166 36 Prague 6, Czech Republic 4 Department of Neurosurgery, Mayo Clinic, Rochester MN, 55905, USA 5 Department of Physiology & Biomedical Engineering, Mayo Clinic, Rochester MN, 55905, USA This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Abstract An ability to map seizure-generating brain tissue, i.e., the seizure onset zone (SOZ), without recording actual seizures could reduce the duration of invasive EEG monitoring for patients with drug-resistant epilepsy. A widely-adopted practice in the literature is to compare the incidence (events/time) of putative pathological electrophysiological biomarkers associated with epileptic brain tissue with the SOZ determined from spontaneous seizures recorded with intracranial EEG, primarily using a single biomarker. Clinical translation of the previous efforts suffers from their inability to generalize across multiple patients because of (a) the inter-patient variability and (b) the temporal variability in the epileptogenic activity. Here, we report an artificial intelligence-based approach for combining multiple interictal electrophysiological biomarkers and their temporal characteristics as a way of accounting for the above barriers and show that it can reliably identify seizure onset zones in a study cohort of 82 patients who underwent evaluation for drug-resistant epilepsy. Our investigation provides evidence that utilizing the complementary information provided by multiple electrophysiological biomarkers and their temporal characteristics can significantly improve the localization potential compared to previously published single-biomarker incidence-based approaches, resulting in an average area under ROC curve (AUC) value of 0.73 in a cohort of 82 patients. Our results also suggest that recording durations between ninety minutes and two hours are sufficient to localize SOZs with accuracies that may prove clinically relevant. The successful validation of our approach on a large cohort of 82 patients warrants future investigation on the feasibility of utilizing intra-operative EEG monitoring and artificial intelligence to localize epileptogenic brain tissue. Broadly, our study demonstrates the use of artificial intelligence coupled with careful feature engineering in augmenting clinical decision making. Keywords Seizure Onset Zone, Epilepsy Surgery, Artificial Intelligence in Neurological Applications, High- Frequency Oscillation, Interictal Epileptiform Discharge, Phase-Amplitude Coupling, and Support Vector Machine. This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Introduction Epilepsy is one of the most prevalent and disabling neurologic diseases. It is characterized by the occurrence of unprovoked seizures and affects ~1% of the world's population [Leonardi 2002]. Many patients with epilepsy achieve seizure control with medication, but approximately one-third of people with epilepsy continue to have seizures despite taking medications [Kwan 2000]. In such cases, one treatment option is surgical resection of the brain tissue responsible for seizures, but this option depends critically on accurate localization of the pathological brain tissue, which is referred to as the seizure onset zone (SOZ). Clinical SOZ localization requires implanting of electrodes for intracranial EEG (iEEG) that is recorded over several days to allow sufficient time for spontaneous seizures to occur [Lüders 2006]. The electrodes that are in the SOZ are identified based on visual inspection of the iEEG captured at the time of seizures, and some tissue around these electrodes is removed during a surgical procedure. Despite being the current gold standard for mapping of the epileptic brain in a clinical setting, this manual procedure is time-consuming, costly, and associated with potential morbidity [Van Gompel 2008, Wellmer 2012]. Recently, the use of interictal (non-seizure) iEEG data for the identification of SOZs and for the possibility of replacing multi-day ICU monitoring to record habitual seizures has received notable interest [van 't Klooster 2015]. A common practice undertaken in the literature investigating interictal SOZ localization is to compare the incidence rates (events/time) of putative pathological electrophysiological events associated with epileptic brain tissue (known as electrophysiological biomarkers of epilepsy) detected in iEEG recorded from individual electrodes against the gold-standard SOZ electrodes determined from spontaneous seizures. Among the potential electrophysiological biomarkers, high-frequency oscillations (HFOs) [Bragin 1999, Bragin 2002, Worrell 2011, Zijlmans 2012] and interictal epileptiform discharges (IEDs) [Staley 2011] have been the most widely investigated. Phase- amplitude coupling (PAC) and other forms of cross-frequency coupling (CFC) have more recently been investigated as promising clinical biomarkers for epilepsy [Papdelis 2016]. HFOs are local field potentials that reflect short-term synchronization of neuronal activity, and they are widely believed to be clinically useful for localization of epileptic brain [Bragin 1999, Worrell 2004, Jirsch 2006, Staba 2002]. Furthermore, there is an extensive literature investigating IEDs as interictal markers of seizure onset zones, but it has met with limited success [Lüders 2006, Marsh 2010, Hauf 2012]. PAC (a measure of cross-frequency coupling) [Jensen 2007] as an adjunct to ictal (seizure) biomarkers was shown to be useful for SOZ localization [Edakawa 2016], and more recently, PAC has been evaluated as an interictal marker for determining SOZs [Amiri 2016]. Most the existing studies have utilized a simple counting of the above biomarkers (detected either manually or using software) in fixed durations to classify the electrodes that are in the SOZ [Cimbalnik 2017]. Although some recent approaches have utilized clustering methods [Liu 2016] and dimensionality reduction methods [Weiss 2016] as preprocessing steps in identifying pathologic interictal HFOs, the determination of SOZs was still performed using simple counting of HFOs. Furthermore, these approaches have predominantly utilized a single biomarker to identify SOZs and have not considered the inter-patient variability nor the temporal dynamics of the epileptic activity [Cimbalnik 2017]. As a result, they have not been able to generalize across multiple patients and their overall accuracies have been insufficient to bring them into clinical practice [Holler 2015, Nonoda 2016, Sinha 2017]. The reasons for inter-patient variability might include electrode placement, false-positive detections of biomarkers, signal artifacts, the varied etiology of focal epilepsy, or the fact that some biomarkers are incident in both physiologic and pathologic states [Matsumoto 2013, Ben-Ari 2007, Cimbalnik 2016]. Thus, utilizing a single biomarker to identify SOZs of patients with potentially heterogeneous epileptogenic mechanisms may result in unsatisfactory accuracy for some individuals. We hypothesize that it may be possible to reduce inter-patient variability by combining the complementary values contained within different electrophysiological biomarkers and thereby improve SOZ localization potential. However, despite the growing interest in each of the above biomarkers, the extent to which they provide independent predictive value for epileptogenic tissue localization remains unclear. From a signal-processing perspective, IED represents a relatively distinct electrophysiological phenomenon compared to HFO and PAC. However, temporal correlations of HFO events and PAC may be observed when short HFOs co-occur with IEDs [Weiss 2016]. Apart from that each biomarker will constitute specific that specific instance, it is possible This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 electrophysiological information about the epileptogenicity of brain tissue and might add predictive value when used in unison with other biomarkers. However, the potential clinical utility of combining electrophysiological biomarkers has received relatively little investigation [Gnatkovsky 2014]. In addition, there is evidence that behavioral states play a role in altering the temporal patterns of epileptiform activity in the brain [Staba 2002, Worrell 2008, Amiri 2016]. As a result, the occurrence of the biomarkers exhibits temporally varying rates when long EEG recordings with mixed behavioral states are considered [Pearce 2013]. Hence, the common practice of simply counting HFOs or IEDs for a fixed duration and using an average rate to determine the SOZ is likely suboptimal. We recently proposed a temporal filtering-based unsupervised approach to utilize temporal characteristics of spectral power features to determine SOZs interictally [Varatharajah 2017]. However, more sophisticated models are needed to effectively utilize multiple electrophysiological biomarkers and their temporal characteristics to accurately determine SOZs. Modern artificial intelligence (AI) based methods facilitate (a) the ability to learn high-dimensional decision functions from labeled training data and (b) the flexibility to define customized features representing domain knowledge [Russell 1995]. We believe that these properties can be useful in harnessing multiple electrophysiological biomarkers and their temporal characteristics for the interictal classification of SOZs. However, AI- based approaches have been underexplored in the SOZ classification literature, mainly because of the unavailability of large-scale iEEG datasets collected during epilepsy-surgery evaluation. Fortunately, the availability of continuous iEEG recordings collected from a large cohort of 82 patients and approximately 5000 electrodes gives us a unique opportunity to assess the potential utility of AI-based approaches in this study. Although this dataset is still not large enough to automatically extract class- specific electrophysiological patterns using deep-learning approaches, comparable performance can be realized on this dataset using careful feature engineering and appropriate model selection. In that context, the aim of this study is to develop an AI-based analytic framework that utilizes multiple interictal electrophysiological biomarkers (e.g., HFO, IED, and PAC) and their temporal characteristics for interictal electrode classification and mapping of SOZs. To that end, we developed a support vector machine (SVM) based classification model utilizing customized features (based on the above biomarkers) extracted from 120-minute interictal iEEG recordings of 82 patients with drug- resistant epilepsy to interictally identify the electrodes representing their seizure onset zones. This approach achieved an average AUC (area under ROC curve) of 0.73 when the HFO, IED, and PAC biomarkers were used jointly, which is 14% better than the AUC achieved by a conventional biomarker incidence-based approach using all three biomarkers, and 4 -- 13% better than that of an SVM-based model that used any one of the biomarkers. This result indicates that exploiting the temporal variations in biomarker activity can improve localization of epileptic brain and that the biomarkers utilized in this study have complementary predictive values. Our analysis of individual patients reveals that the AUCs improve or remain unchanged for more than 65% of the patients when the composite, rather than any single biomarker, is utilized, supporting the hypothesis that combining multiple biomarkers can provide more generalizability than can individual biomarkers. Development of this technology also enabled us to develop an understanding of the recording durations required for interictal localization of SOZs. By analyzing iEEG segments of different durations (10 -- 120 minutes), we show that longer iEEG segments provide better accuracy than do short segments, and that the improvements become statistically insignificant for durations beyond 90 minutes. These promising results warrant further investigation on the feasibility of using intra-operative mapping to localize epileptic brain. This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Methods Experimental Setup Data used in this study were recorded from patients undergoing evaluation for epilepsy surgery at the Mayo Clinic, Rochester, MN. The Mayo Clinic Institutional Review Board approved this study, and all subjects provided informed consent. Subjects underwent intracranial depth electrode implantation as part of their evaluation for epilepsy surgery whenever noninvasive studies could not adequately localize the SOZ. To provide an unbiased dataset for analysis, we took 2 hours of continuous iEEG data (during some period between 12:00 and 3:00 AM) on the night after surgery. Subjects Data from 82 subjects (48 males and 34 females, with an average age of 31) with focal epilepsy were investigated by post hoc analysis. All subjects were implanted with intracranial depth arrays, grids, and/or strips; see supplementary Table 1 for details. Subjects underwent multiple days of iEEG and video monitoring to record their habitual seizures. Electrodes and Anatomical Localization Depth electrode arrays (from AD-Tech Medical Inc., Racine, WI) were 4- and 8-contact electrode arrays consisting of a 1.3-mm-diameter polyurethane shaft with platinum/iridium (Pt/Ir) macroelectrode contacts. Each contact was 2.3 mm long, with 10-mm or 5-mm center-to-center spacing (with a surface area of 9.4 mm2 and an impedance of 200 -- 500 Ohms). Grid and strip electrodes had 2.5-mm-diameter exposed surfaces and 1-cm center-to-center spacing of adjacent contacts. Anatomical localization of electrodes was achieved using post-implant CT data co-registered to the patient's MRI using normalized mutual information [Ashburner 2008]. Electrode coordinates were then automatically labeled by the SPM Anatomy toolbox, with an estimated accuracy of 0.5 mm [Tzourio-Mazoyer, et al., 2002]. Signal Recordings All iEEG data were acquired with a common reference using a Neuralynx Cheetah electrophysiology system. (It had a 9-kHz antialiasing analog filter, and was digitized at a 32-kHz sampling rate, filtered by a low-pass, zero-phase-shift, 1-kHz, low-pass Bartlett-Hanning window, and down-sampled to 5 kHz.) Clinical SOZ Localization The SOZ electrodes and time of seizure onset were determined by visually identifying the electrodes with the earliest iEEG seizure discharges. Seizure onset times and zones were determined by visual identification of a clear electrographic seizure discharge, followed by looking back at earlier iEEG recordings for the earliest electroencephalographic change contiguously associated with the visually definitive seizure discharge. The same approach has been used previously to identify neocortical SOZs [Worrell 2004] and medial temporal lobe seizures [Worrell 2008]. The identified SOZ electrodes were used as the gold standard to test and validate our analyses. Data Continuous 2-hour interictal segments of iEEG data, sufficiently separated from seizures, were chosen for all 82 patients to represent a monitoring duration that could be achieved during surgery. A total of 4966 electrodes were implanted across the 82 subjects, and 911 of them were identified to be in SOZs via ictal localization performed by clinical epileptologists caring for the patients. Data Preprocessing Prior to analysis, continuous scalp and intracranial EEG recordings were reviewed using a custom MATLAB viewer [Brinkmann 2009]. Electrode channels and time segments containing significant This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 artifacts or seizures were not included in subsequent analyses. All iEEG recordings were filtered to remove 60-Hz power-line artifacts. Overall Analytic Scheme Selected 2-hour iEEG recordings were divided into non-overlapping 3-second epochs. A 3-second epoch length was chosen to accommodate at least a single transient electrophysiologic event (in the form of a PAC, HFO, or IED) that could be associated with the SOZ. The HFO [Cimbalnik 2017], IED [Barkmeier 2012], and PAC [Amiri 2016] biomarkers were extracted using previously published detectors to measure their presence in each 3-second epoch. Then, a clustering procedure was performed to assign a binary observation of normal or abnormal to each channel. This procedure was performed separately for each patient, and the biomarker measures extracted in a 3-second recording of all the channels of a specific patient were considered. These channels were clustered into two groups based on their similarities with respect to each biomarker, and the cluster with the larger average biomarker rate was considered the abnormal cluster. This step was performed so that biomarker detections that had strong magnitudes and showed strong spatial correlation were retained, and electrodes with noisy detections were minimized. At the end of that procedure, every 3-second recording of a channel was associated with three binary values (one for each biomarker) representing the presence of the HFO, IED, and PAC biomarkers. We refer to those binary values as observations. Since there are 2400 3-second epochs in a 2-hour period, the total number of observations made in a channel was 3 * 2400 = 7200. Since that number of features is relatively large compared to the number of channels available in our study, we reduced the number of observations by applying a straightforward summation approach. Binary observations made within a 10-minute window (200 epochs) were counted to arrive at a measure of the local rate of the biomarker incidence for each 10- minute window of the 2-hour recording. Although the 10-minute window length may appear to have been chosen arbitrarily, it is large enough to have less noisy local biomarker incidence rates and yet not so large as to mask the temporal variations in interictal biomarker activity. This method reduces the number of observations for a channel to 36 (3 local biomarker rates × 12 windows). These observations, made across a 2-hour period of a channel, were used to infer whether that channel belonged to an SOZ under a supervised learning setting using a support vector machine (SVM) classifier. The whole process is illustrated as a flow diagram in Figure 1. This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Figure 1: The overall analytic scheme of the SOZ detection algorithm utilized in this study. A 2- hour data segment is analyzed for each patient. PAC, HFO, and IED biomarkers are extracted in 3- second epochs, and a clustering method is used to group channels based on similarities with respect to the biomarkers. These groupings are converted to binary (0, 1) observations and counted within a 10- minute window to obtain local biomarker incidence rates. These local biomarker incidence rates for the three biomarkers within all the 10-minute windows of a channel's 2-hour recording are utilized as the features of that channel in a machine learning setting. A support vector machine (SVM) classifier, which was trained and tested using labeled training data, is used to predict whether an electrode is in an SOZ. Detection of Interictal Electrophysiological Biomarkers The PAC measure was calculated by correlating instantaneous phase of the low-frequency signal with the corresponding amplitude of a high-frequency signal for a given set of low- and high-frequency bands (Figure 2a). In this implementation, low- and high-frequency contents in the signal were extracted using MORLET wavelet filters, and all frequency bands were correlated against all others to create a so-called PAC-gram (Figure 2b). Based on the observed high PAC content and the existing literature [Weiss 2015], 0.1 -- 30 Hz was chosen as the low-frequency (modulating) signal, and 65 -- 115 Hz was chosen as the high-frequency (modulated) signal in the rest of the analysis. HFOs were detected using a Hilbert transform-based method [Kucewicz & Berry 2015, Pail 2017, Cimbalnik 2016]. The data segments were bandpass-filtered for every 1-Hz band step from 50 to 500 Hz. Then, the filtered-data frequency bands were normalized (z-score), and the segments in which the signal amplitudes were three standard deviations above the mean for a duration of one complete cycle of a respective high frequency (in 65 -- 500 Hz) were marked as HFOs (see Figure 2c) [Matsumoto 2013]. IEDs were extracted using a previously validated spike-detection algorithm [Barkmeier 2012]. A detection threshold of four standard deviations (of differential amplitude) around the mean was used to mark IEDs in this algorithm (see Figure 2d). The HFO, IED, and PAC detected events were stored in a database. This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Figure 2: Phase amplitude coupling (PAC), high-frequency oscillations (HFO), and interictal epileptiform discharge (IED) detection. (A) Detailed illustration of the PAC feature extraction algorithm. Low (0.1 -- 30 Hz) and high (65 -- 115 Hz) frequency components are filtered out from the raw signal. The phase of the slow wave is correlated with the high-frequency amplitude envelope to measure coupling. (B) A PAC-gram representing the average interictal PAC measured between different frequency bands. Highlighted portion indicates the low- and high-frequency bands utilized in the rest of our analysis. (C) Pictorial illustration of HFO detection. Oscillations that have an amplitude of three standard deviations above the mean and lasting for more than one complete cycle in low- gamma (30 -- 60 Hz), high-gamma (60 -- 100 Hz), and ripple (100 -- 150 Hz) bands are detected. (D) An illustration of detected IEDs. Differential amplitude is standardized, and a threshold of four standard deviations around the mean was used to mark IEDs. Prediction of SOZ Electrodes Using a Support Vector Machine Classifier The biomarkers extracted from a 2-hour recording of a channel were converted to a 36-dimensional feature vector as shown in Figure 1. The features represent the local biomarker incidence rates (with a separate rate for each biomarker) within each 10-minute window of the 2-hour recording. These features were standardized to eliminate any differences in scale. We took two different approaches to perform cross-validation. First, we performed 10-fold cross validation. The dataset, including standardized features of 4966 electrodes (including 911 SOZ electrodes), was divided into a 60% training set and a 40% testing set, keeping the same proportion of SOZ and NSOZ (non-SOZ) electrodes in both sets. Second, we performed leave-1-out cross-validation. For every subject in the dataset, we used the data from the rest of the subjects as the training data and the respective subject's data as the testing data. In each of the cross-validation iteration, training and testing datasets were generated using one of the cross-validation approaches. An SVM classifier was trained on the training set, whose hyper-parameters (described below) we optimized by performing a grid search with a tenfold cross-validation within the training set. The classifier trained on the best-performing hyper- This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 parameters was used to predict the labels of the channels in the testing set. By comparing those predictions against the ground-truth labels of the testing set channels, we calculated the metrics of model fitness. This process was repeated 10 times with different combinations of training and testing sets to obtain metrics of generalized performance. This process is illustrated in Figure 3. Figure 3: A flow diagram illustrating the prediction framework. The input is the whole dataset, including the 36 features extracted from 82 subjects (4966 channels) and their gold-standard labels assigned by clinical epileptologists. (A) 10-fold cross-validation: first, the dataset is shuffled to randomize the order of channels in it. After randomization, the dataset is partitioned into two sets, a 60% training set and a 40% testing set, keeping the same proportion of NSOZ and SOZ channels in both sets. (C) Inner CV loop: A set of optimal hyper-parameters is selected for the SVM classifier based on a tenfold cross-validation within the training set. (D) Goodness-of-fit metrics: The classifier learned in the previous step is tested on the testing dataset, and measures of its performance are generated. This whole procedure is repeated 10 times (i.e., 10-fold CV) or 82 times (i.e., leave-1-out CV) to produce generalized performance metrics, eliminating any bias introduced by a specific split of training and testing sets. A support vector machine is a binary classifier that finds the maximum margin hyper-plane that separates the two classes in the data [Boser 1992]. The data being classified are denoted by 𝑋 ∈ ℝ𝑁×𝑃 (𝑁 channels and 𝑃 features), and the data from channel 𝑖 are denoted by 𝑋(𝑖) ∈ ℝ𝑃. The class labels for all the channels are denoted by 𝑌 ∈ {−1,1}𝑁 (where −1 and 1 are numerical labels for the two classes), and the class label for channel 𝑖 is denoted by 𝑌(𝑖) ∈ {−1,1}. The optimization problem to This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 find the optimal hyper-plane (described by weights 𝑊 ∈ ℝ𝑃 and intercept term 𝑏 ∈ ℝ) is shown in Eq. 1. min W,b 1 2 ‖𝑊‖2 (1) subject to 𝑌(𝑖)(𝑊 𝑇𝑋(𝑖) + 𝑏) ≥ 1, 𝑖 ∈ {1, … , 𝑁} Once the optimal hyper-plane [𝑊𝑜𝑝𝑡, 𝑏𝑜𝑝𝑡] is found, the predicted class label for channel 𝑖 is obtained 𝑇 𝑋(𝑖) + 𝑏𝑜𝑝𝑡. This formulation assumes that the data have a clear separation as the sign of 𝑊𝑜𝑝𝑡 between the two classes. When that is not the case, slack variables and a tolerance parameter (box- constraint) can be introduced to obtain separating hyper-planes that tolerate small misclassification errors [Cortes 1995]. Dual formulation of SVM has received considerable interest because it enables use of different kernel transformations of the original feature space without altering the optimization task and because of its advantages in complexity when the data are high-dimensional [Boser 1992]. Specifically, this allows for the features to be transformed from the original feature space to a kernel space. With this transformation, the cases in which the original data are not linearly separable may be solved because transformation of the data to higher dimensions may introduce linear separation in the transformed domain. Linear, radial basis function (RBF), and polynomial kernels are widely used kernels in this context. Goodness of fit of the SVM classifier is evaluated by predicting the classes of the test dataset by using the classifier that was trained on the training dataset and comparing the predictions against the true class labels of the test dataset. This comparison is performed using standard performance metrics such as receiver operating characteristics (ROC) curve analysis, area under ROC curve (AUC), sensitivity, specificity, accuracy, precision, recall, and F1-score. Because of heterogeneity in the data, choosing one partition of training and test datasets is not sufficient to credibly evaluate the performance of a classifier. A common practice to obviate the effect of heterogeneity in the data is to perform several iterations of training-testing cross-validation of the dataset. One run of this procedure is carried out by choosing a subset of the dataset as training data, and testing on the rest of the dataset. This approach allows the calculation of generalizable performance metrics for the analyzed classifier. Results Nonlinear Classification Boundary between SOZ and NSOZ Electrodes Understanding the nature of the separation between the two classes in feature space is important to achieve the maximum classification performance in binary classification. If the separation is linear, a linear classifier should be sufficient (and preferable due to the Occam's razor principle) to achieve the maximum attainable classification performance. On the other hand, when the separation is nonlinear, linear classifiers perform poorly compared to nonlinear classifiers. However, when the feature space is high-dimensional, visualizing the boundary between classes can be difficult. An option is to use linear and nonlinear classifiers to classify the two classes and plot the histograms of likelihood probabilities predicted by the classifier to understand the degree of separation achieved by linear and nonlinear boundaries [Cherkassky 2010]. We performed this analysis for our dataset by using an SVM classifier with linear and RBF kernels. Furthermore, we used the 10-fold cross-validation approach to perform this analysis because the individual AUCs obtained using the leave-1-out cross-validation approach were highly variable across patients. We trained two SVM classifiers with linear and RBF kernels, respectively, using the framework shown in Figure 3 with 60% of all the electrodes and all the biomarkers. These classifiers were used to predict the class labels for the rest of the electrodes (40%). We then compared the distribution of the likelihood probabilities generated by the two SVM classifiers against the true class labels. Figures 4a and 4b show the histograms of the likelihood probabilities obtained using an SVM classifier with linear and RBF kernels, respectively, for SOZ and NSOZ electrodes in the testing set. The nonlinear boundary achieved through an SVM classifier with This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 an RBF kernel clearly has a better separation between the two classes, as can also be seen in Figure 4c. To quantify this observation, we plotted ROC curves (which are shown in Figure 4d) for the predictions obtained using linear and RBF kernels. The linear-SVM classifier obtained an AUC value of 0.57, while the RBF-SVM classifier obtained an AUC of 0.79 when all the biomarkers were utilized. Hence, we conclude that the boundary between SOZ and NSOZ electrodes is nonlinear in the feature space in which the features are derived as described in Figure 1, and the rest of our analyses focus on the results obtained by the SVM classifier with an RBF kernel. Figure 4: Results obtained using a support vector machine with interictal electrophysiological biomarkers to classify seizure onset zone (SOZ) electrodes. (A) and (B) show the probability densities of the likelihoods predicted by an SVM classifier for SOZ and NSOZ electrodes in the testing set for linear and RBF kernels, respectively. The RBF kernel results in less overlap between the SOZ and NSOZ probability densities. (C) Boxplot showing the range of likelihood probabilities obtained for SOZ and NSOZ electrodes when all biomarkers were used as features in a n SVM classifier with an RBF kernel. (D) A comparison between the ROC curves when an SVM classifier was used with an RBF kernel for individual biomarkers and their combination, and when it was used with a linear kernel with a combination of all biomarkers. (E) A comparison between the AUCs obtained using conventional unsupervised methods that use overall rates of biomarker incidence to predict SOZ electrodes, and those obtained using our SVM-based supervised approach. A Supervised Learning Approach Improves the SOZ Localization Accuracy This paper proposes a supervised-learning-based approach that uses an SVM classifier to predict electrodes in an SOZ by using interictal iEEG data. In order to understand whether our supervised approach is better than simply using biomarker rates, we implemented biomarker-rate-based SOZ This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 electrode classification for HFO, IED, and PAC separately and in combination. (We simply added the individual biomarker incidence rates to obtain an overall biomarker incidence rate.) Figure 4e shows a comparison between the ROC curves obtained for the unsupervised biomarker incidence rate-based approach and the supervised approach that uses an SVM classifier with an RBF kernel when all biomarkers were utilized. Predictions using the unsupervised approach were performed on the same testing set electrodes that were used in the supervised approach. The SVM-based supervised approach outperformed the unsupervised approaches with a 17 -- 23% gain in the AUC value. Notably, the performance of the SVM classifier with a linear kernel was comparable to that of the unsupervised approach, with an AUC value of 0.57, as seen in Figure 4d. This highlights the ability to significantly improve the correct classification of previously unseen SOZ electrodes by utilizing the right machine learning method (in this case, SVM with an RBF kernel) to learn the characteristics of SOZ electrodes. Goodness-of-Fit Metrics for SOZ Electrode Classification Other goodness-of-fit metrics, as specified previously, were calculated for the different combinations of biomarkers and methods and are listed in Table 1. For all the metrics other than AUC, the likelihood probabilities assigned by the classifier were applied with a threshold to classify SOZ and NSOZ electrodes. In order to compare the different approaches, this threshold was chosen using a common criterion, i.e., the false positive rate is approximately 25%. AUCs obtained for the training set are reported in supplementary Table 3 along with optimal hyper-parameters used in each of the cross-validations. Table 1: Cross-validated goodness-of-fit metrics for SOZ determination. Here we list the goodness-of-fit metrics (AUC, sensitivity, specificity, accuracy, precision, recall, and F1-score) obtained for the test dataset, for the different combinations of biomarkers and analytic techniques shown in Figure 4. The average values and standard deviations were computed using a) a tenfold stratified cross-validation and b) a leave-one-out cross-validation. Bio- marker 10-fold CV All All Method AUC Sensitivity Specificity Accuracy Precision (%) (%) (%) (%) Recall (%) F1-score (%) SVM-LIN 0.56(0.03) 32.20(4.17) 75.09(0.02) 67.23(0.78) 22.42(2.31) 32.2(4.17) 26.43(3.01) SVM-RBF 0.79(0.01) 70.36(1.78) 75.09(0.00) 74.22(0.33) 38.79(0.60) 70.36(1.78) 50.01(0.95) HFO SVM-RBF 0.68(0.01) 53.71(1.70) 75.16(0.09) 71.23(0.29) 32.66(0.66) 53.71(1.70) 40.62(1.00) IED PAC ALL HFO IED PAC SVM-RBF 0.68(0.01) 55.11(2.53) 75.07(0.03) 71.41(0.45) 33.14(1.01) 55.11(2.53) 41.39(1.50) SVM-RBF 0.73(0.01) 60.63(2.74) 75.09(0.02) 72.44(0.51) 35.31(1.03) 60.63(2.74) 44.63(1.57) RATE 0.58(0.01) 35.91(2.03) 75.09(0.02) 67.91(0.37) 24.43(1.03) 35.91(2.03) 29.07(1.40) RATE 0.56(0.01) 32.39(2.20) 76.31(0.62) 68.26(0.77) 23.48(1.46) 32.39(2.20) 27.22(1.74) RATE 0.58(0.01) 35.19(1.42) 75.18(0.09) 67.85(0.27) 24.14(0.74) 35.19(1.42) 28.63(0.99) RATE 0.62(0.01) 43.76(2.25) 75.27(0.12) 69.49(0.42) 28.41(1.03) 43.76(2.25) 34.45(1.46) Leave one (subject) out CV SVM-RBF 0.73(0.02) 57.45(2.82) 79.49(0.57) 73.3(0.90) 38.71(2.45) 57.45(2.82) 43.49(1.99) SVM-RBF 0.63(0.01) 35.53(2.89) 84.16(0.86) 73.10(1.00) 34.63(2.72) 35.53(2.89) 35.09(2.11) SVM-RBF 0.60(0.01) 33.50(2.22) 80.77(0.66) 69.47(0.89) 30.55(2.42) 33.50(2.22) 29.16(1.55) SVM-RBF 0.69(0.01) 47.70(2.81) 81.03(0.63) 72.63(0.93) 36.23(2.39) 47.70(2.81) 39.06(1.94) ALL HFO IED PAC This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Combining Multiple Electrophysiological Biomarkers Improves the Localization Accuracy The classification framework depicted in Figure 3 was utilized with the features relevant to HFO, IED, and PAC biomarkers separately to reveal the predictive ability of individual biomarkers. Then the individual classification performances were compared against the performance obtained when all the biomarkers were used together. Figure 4d shows the ROC curves obtained for the different runs of the classification framework. While the PAC biomarker had the best predictive ability individually (AUC: 0.74 -- 10-fold CV, 0.69 -- leave-1-out CV), the classification obtained using all the biomarkers together performed better than any of the individual biomarkers, providing an AUC of 0.79 with the 10-fold CV approach and an AUC of 0.73 with the leave-1-out CV approach. These findings support the idea that the different interictal electrophysiological biomarkers used in this study possess complementary information that can be harnessed to achieve a superior performance in predicting the electrodes in an SOZ. Figure 5: Improvements in patient-specific SOZ classification achieved by means of combining multiple biomarkers as opposed to utilizing a single biomarker in the SVM framework. (A), (B), and (C): Improvements obtained in AUCs for patient-specific SOZ classification when the combination of multiple biomarkers was utilized compared to when HFO, IED, or PAC, respectively, was utilized alone. (D) Histogram densities of the patient-specific AUCs for the prediction of SOZ electrodes using HFO, IED, PAC, and their composite. To determine whether the combination of multiple biomarkers can reduce the inter-patient variability, we analyzed the improvements in SOZ electrode classification potential in each individual by calculating the AUC for each individual separately. We predicted the SOZ electrodes of each This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 individual patient using HFO features, IED features, and PAC features separately and then their composite, within the SVM-based framework described previously using a leave-one-out cross- validation approach (see supplementary Table 2). Figures 5a -- 5c illustrate the respective improvements in the AUCs of individual patients attained when the combination of the three biomarkers was utilized instead of HFO, IED, or PAC by itself. Our analysis indicates that the AUCs of individual patients improved or remained unchanged for more than 65% of the patients when the composite was utilized compared to any individual biomarker. Then we plotted the histograms of the AUCs of individual patients for each biomarker and their composite separately, and approximated the densities by using kernel density estimation. Figure 5d shows that the histogram density of AUCs of individual patients becomes skewed towards higher AUC values when the composite is utilized to predict SOZ electrodes. This indicates that the utilization of multiple biomarkers with complementary information reduces the overall variability across patients in the ability to classify SOZ electrodes -- variability that is apparent with any single biomarker. Recording Durations Between 90 and 120 Minutes May be Sufficient for Interictal SOZ Localization The ability to localize seizure-generating brain tissue is a cornerstone of clinical epileptology. We investigated how the duration of interictal iEEG recording impacted the the localization of the SOZ (as illustrated in Figure 6a). Here we applied our AI-based framework on a range of recording durations between 10 and 120 minutes. Figure 6 shows the mean ROC curves obtained using a tenfold cross-validation for different durations when an SVM classifier with an RBF kernel was utilized with all the biomarkers. To quantify the different runs, we plotted the AUC metric against the recording length used for SOZ electrode classification prediction. That is shown in Figure 6c, where the error bars indicate the standard deviations of the AUCs based on a tenfold cross-validation. Statistical significance tests using two-tailed paired t-tests indicate that the AUCs obtained using 90-, 100-, and 110-minute recordings are not statistically very different from the AUCs obtained using a 120-minute recording. This finding indicates that recording durations between 90 and 120 minutes may be sufficient for interictal SOZ identification with clinically relevant accuracies. Figure 6: Evaluation of the length of recordings and interictal SOZ localization. (A) ROC curves obtained when shorter interictal segments of durations ranging from 20 to 120 minutes were utilized for analysis. (B) AUC values obtained with short interictal segments. Longer interictal segments result in better AUC values; however, the AUCs obtained using segments longer than 90 minutes are statistically indifferent (based on 2-sided paired t-tests between AUCs). This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 20406080100120Recording duration (minutes)0.60.650.70.750.80.85Area Under ROC Curve (AUC)00.20.40.60.81False positive rate00.20.40.60.81True positive rate20 minutes40 minutes60 minutes80 minutes100 minutes120 minutes(A)(B)NSNSNSNS: P<0.05: Not significant Discussion Main Contribution of the Study The current study describes a machine learning method for classification of SOZ and NSOZ electrodes using multiple electrophysiological biomarkers extracted from interictal iEEG data collected in a clinical setting. Our study, to our knowledge, is the first to utilize the complementary information provided by multiple electrophysiological biomarkers and their temporal characteristics as a way of reducing variability across patients to improve interictal SOZ localization. Using 2-hour wide-bandwidth intracranial EEG recordings of ~5000 electrodes from 82 patients, our study provides a large-scale evaluation of an artificial intelligence-based approach for interictal SOZ localization and shows that when used in concert with multiple interictal biomarkers, it can outperform single biomarker approaches. The ability to perform real-time feature processing and SOZ determination with clinically relevant accuracies, with a maximum monitoring duration of 2 hours, supports the feasibility of SOZ determination using interictal intracranial EEG data (see supplementary Table 2 for specific computation times for each task included in interictal SOZ classification). In addition to the above specific contributions, our study also exemplifies the role of artificial intelligence in augmenting clinical workflows. The Value of Combining Multiple Biomarkers Interictal SOZ localization techniques have been widely discussed, with the main focus being on the search for and validation of a single biomarker that can be used in all patients [Bragin 1999, Jacobs 2010, Jacobs 2008, Worrell 2011, Engel 2013]. Conventional methods have focused on HFO biomarker detection algorithms and on HFOs themselves [Worrell 2012, Burnos 2014, Balach 2014]. However, the generalizability of such single biomarkers has been insufficient for clinical practice [Nonoda 2016, Sinha 2017, Holler 2015, Cimbalnik 2017], primarily because of inter-patient variability, and it appears that one biomarker may not be sufficient to identify SOZs in all patients. While there have been multiple attempts to automate SOZ localization [Liu 2016, Graef 2013, Gritsch 2011], very little work has attempted to improve localization potential by means of combining multiple biomarkers. This exploratory study shows interictal electrophysiological biomarkers within a rigorous, supervised machine learning setting can be more accurate in performing interictal SOZ localization than can utilization of a single biomarker, essentially by reducing inter-patient variability. This is evident from our individual patient-based analysis (Figure 5), in which we show that SOZ electrode classification AUCs improve or remain unchanged for more than 65% of patients when the combination of the three biomarkers is utilized instead of single biomarkers. This finding indicates that combining multiple biomarkers reduces the variance in SOZ electrode classification and therefore achieves better generalizability than single- biomarker-based approaches. Figure 5 also shows that combining multiple biomarkers reduces the accuracy in SOZ electrode classification for some patients when compared with any of the individual biomarkers. The reason for this could be that there are more disagreements within the different biomarkers than agreements with respect to SOZ electrodes. The situations in which combining multiple biomarkers is detrimental can be explored and tested by comparing the individual predictions provided by the biomarkers. that combining multiple Exploiting the Temporal Variability in Epileptic Activity to Improve Classification Potential We showed in this study that utilizing a machine learning approach that uses local rates of biomarkers within 10-minute subintervals for classifying SOZ electrodes improves the AUC by 19% compared to traditional unsupervised approaches that primarily utilize the overall rates of biomarkers. We believe that the improvement is due to the inability of the overall-rate-based approaches to account for the temporal variations in epileptic activity. Recent studies have reported that the rate of epileptiform activity changes significantly between different behavioral states [Worrell 2008, Amiri 2016]. Since our study uses night segments with mixed behavioral states, the ability to differentiate SOZ electrodes from NSOZ electrodes using rates of epileptiform activity varies depending on the This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 portion of the segment. Hence, looking at the overall rates of epileptiform activity might average out the variations between subintervals and hence degrade performance, and result in accuracy lower than the maximum attainable. We also show that a nonlinear classification technique provides significant improvements in AUC compared to a linear classification technique and that the performance provided by the latter is similar to that of commonly used unsupervised approaches. Although a linear classifier considers each subinterval for classification, it is impractical to assign a particular subinterval a higher weight because the exact subintervals in which the epileptiform activity is highly discriminative may not be the same across different patients. Therefore, when trained across multiple patients, the linear classifier perceives each subinterval as equally important and assigns all of them equal weights. As a result, its performance is similar to that of the approaches that use overall epileptiform activity rates to classify SOZ electrodes. On the other hand, an RBF kernel measures similarities (Euclidean distances) between a channel and selected SOZ and NSOZ channels (known as support vectors) with respect to their local rates of epileptiform activities. Therefore, regardless of the position of the subinterval that is highly discriminative between SOZ and NSOZ channels, that discriminative ability will be reflected in the overall distance between those channels. That is pictorially illustrated in Figures 7a -- 7d using an example. Figure 7a shows that there is a large variation in the local rates of PAC (in 10-minute windows) across channels of a selected patient. Figure 7b shows the overall PAC rates for each channel, and shows that classifying SOZ electrodes simply by thresholding the overall rates results in poor sensitivity and specificity. Figure 7c shows that transforming the local rates to a kernel space using a linear kernel does not produce any major changes compared to the previous approach with regard to efficiency (in terms of sensitivity and specificity), whereas Figure 7d shows that transforming the local rates using an RBF kernel produces a more favorable transformation because it provides improved sensitivity and specificity. As shown in Figures 7c and 7d, kernel distances between feature vectors of individual channels and average feature vectors of the SOZ channels were calculated using linear and RBF kernels. This example elucidates the utility of a nonlinear machine learning classification approach and its relationship to the underlying physiology. It is also noteworthy that a nonlinear classification approach is suitable in this case due to the manner in which the features of channels were derived, and that linear classifiers may be sufficient for a different derivation of the features. This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Figure 7: Temporal variability in the rates of epileptiform activity (with respect to PAC) and its relation to nonlinear classification. This figure illustrates that the RBF kernel is more specific in capturing the similarities between electrodes with regards to their epileptiform activity patterns. (A) Normalized local PAC rates in different 10-minute intervals of SOZ and NSOZ channels of a selected patient. (B) Overall PAC rates (obtained by summing local rates) for the channels and their poor ability to classify SOZ electrodes. (SS is sensitivity, SP is specificity, and a dashed line indicates application of a threshold). (C) Features after application of a transformation using the linear kernel. This transformation does not provide any notable improvement compared to the summation approach. (D) Features after application of a nonlinear transformation using the RBF kernel. It is evident that this transformation provides better discrimination between SOZ and NSOZ channels. Identifying SOZs During Long and Short Recordings of Night Segments In this study, we showed that localizing an SOZ using multiple features identified in 120 minutes of mixed behavioral state data provided accuracy similar to that obtained with shorter segments (see Figure 6). There is a clear relationship between the utility of this platform and the time of the recording used in analysis. Segments of 120 minutes were arbitrarily chosen to be the maximum amount of time a neurologist could use during an operating room recording to identify the SOZ. Interestingly, results as measured by the AUC did not significantly differ for recording durations between 90 and 120 minutes. It appears that, given satisfactory recording conditions, less than one hour may be all that is needed to achieve high SOZ identification accuracy with this platform. The relationship between the sleep-wake cycle and epileptiform discharges is well known [Sammaritano 1991, Staba 2002, Malow 1998]. Investigation of PAC, HFO, and IED using both short-term and long-term iEEG data shows an increment of the HFO rate with the non-rapid eye This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 movement (NREM) sleep stage, especially in subjects with temporal lobe epilepsy [von Ellenrieder 2017, Amiri 2016, Staba 2004] [Engel 2009]. In addition, recent work has shown that the PAC localization potential increases in slow-wave sleep [Amiri 2016]. Our results suggest that obtaining sleep recordings of a sufficient duration in the clinical routine can be beneficial to quantitative SOZ localization. Contributions Towards Advancing the Current State of Clinical Decision-Making We demonstrated that our approach can identify electrodes in an SOZ with an accuracy of approximately 80% (AUC) on a large cohort of 82 patients, using interictal iEEG recordings of durations less than two hours. The idea of utilizing artificial intelligence to augment clinical workflows has become central in the era of "big data." Studies have shown the utility of deep- learning-based approaches in augmenting clinical diagnosis of skin cancer and diabetic retinopathy, primarily using imaging measurements [Esteva 2017, Gulshan 2016]. These studies have benefited considerably from the availability of a) huge image databases and b) substantially validated artificial neural-network-based classification models. Our study, on the other hand, embodies an alternative approach that utilizes feature engineering as a remedial option for the unavailability of large datasets at the scale of the currently available imaging datasets (with millions of medical images). We believe that the insights drawn from this study will be particularly useful for tasks that lack an abundance of labeled training samples, which inevitably is the case for most clinical problems. Study Limitations Our approach in its current implementation does not possess the ability to differentiate pathological and physiological electrophysiological events. For instance, HFOs are also associated with normal physiological function, and how to distinguish physiological HFOs from pathological HFOs is an active research area [Matsumoto 2013]. Recent studies have shown that utilizing only the pathological events results in increased sensitivity and specificity in determining SOZs interictally [Weiss 2016]. Hence, the ability to differentiate pathological events and then utilize them in determining SOZs may further improve our localization accuracy. As shown recently [Amiri 2016], there is a clear connection between behavioral/sleep state and each of the biomarkers implemented in this study. We also showed that the temporal variations in epileptiform activity due to changes in behavioral states influences the machine learning paradigm utilized in this work. Therefore, accurate annotations of behavioral states considered together with interictal electrophysiologic biomarkers may further improve classification of epileptic and normal brain tissue. However, the patients in this cohort did not have scalp EEGs as required for accurate behavioral state classifications. New methods that can classify sleep stages based on intracranial recordings could prove useful for future analyses [Kremen & Duque 2016]. A successful clinical translation of this approach would depend on the accuracy of this approach in the data collected under operative settings. Prior studies evaluating IEDs [Wass 2001, Schwartz 1997] and HFOs [Zijlmans 2012, Wu 2010] for their potential to localize epileptic brain under intraoperative settings show promise. The PAC biomarker, however, despite being the best individual predictor in interictal settings, has not been evaluated under intraoperative settings and its specificity in such settings is still unclear. Hence, future studies evaluating multiple epilepsy patients are required to accurately determine the clinical utility of this approach. Another limitation of our approach is that we take clinical SOZ as the gold standard in determining the accuracy of the interictal approach. This is a limitation because only a fraction of the patients going through epilepsy patients achieve complete seizure freedom (i.e., ILAE outcome 1). However, the goal of this study is to evaluate how accurately interictal biomarkers can localize the ictal recording SOZ localization. We make the assumption that any clues regarding SOZs that were not captured in ictal localization are less likely to be captured using interictal localization. Regardless, we observed that there is a good agreement between ictal and interictal SOZ localization approaches when the patients eventually had good outcomes (see supplementary Figure 1). This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Future Directions Advances in neuroimaging methods have advanced non-invasive seizure localization capabilities in epilepsy. Because of the holistic spatial view made possible by imaging techniques, they can provide independent information about potential regions of epileptogenic brain that could be used in concert with interictal electrophysiological biomarkers to further improve localization of pathological brain tissue. Alternatively, one can see the utility of combining pathological event classification, behavioral state classification, and multiple imaging modalities as another way of accounting for inter-patient variability, in addition to utilizing multiple interictal electrophysiological biomarkers. However, combining such different data types might require a more complex analytic technique in order to effectively capture the complementary information provided by each data type. We hypothesize that probabilistic graphical models provide an exceptional platform for handling such complexities, as shown in a recent work that combined spatial and temporal relationships in EEGs using a factor-graph-based model [Varatharajah 2017]. To that end, our future work will be directed towards harnessing the utilities of probabilistic graphical models in combining the forenamed multimodal data. Data and software availability The iEEG data and the software used in this study for AI-based interictal SOZ identification is available for download at ftp://msel.mayo.edu/EEG_Data/ and ftp://msel.mayo.edu/JNE_bundle.zip respectively. Conclusions Current methods for localizing seizure onset zones (SOZs) in drug-resistant epilepsy patients are lengthy, costly, and associated with patient discomfort and potential complications. Furthermore, among patients undergoing surgical resection for treatment of drug-resistant epilepsy, approximately 46% suffer seizure recurrence within 5 years. The current gold standard for localizing the tissue to be surgically resected is visual review of ictal (seizure) recordings. In this paper, we reported an artificial intelligence (AI) based approach for the prediction of SOZ electrodes using only non-seizure data. This technique was validated using interictal iEEG data clinically collected from 82 patients with drug-resistant epilepsy. The approach uses three pathological electrophysiological transients reported to be interictal (non-seizure) biomarkers of epileptic brain tissue: high-frequency oscillations (HFOs), interictal epileptiform discharges (IEDs), and phase-amplitude coupling (PAC). We utilize the complementary information provided by these 3 biomarkers and their temporal dynamics in a support vector machine classification (SVM) paradigm to reduce variability across patients and eventually to achieve an average area under ROC curve (AUC) measure of 0.73 in correctly classifying SOZ electrodes interictally. Furthermore, our results suggest that recording durations of approximately two hours are sufficient to localize the SOZ. Some potential applications of this technology are intra- operative electrode placement for prolonged monitoring, electrode placement for electrical stimulation devices [Fisher 2014] to treat seizures, and guiding of the margins of resection around epileptic structural lesions seen on MRI (magnetic resonance imaging). The potential to reduce the number of patients requiring long-term (multiple-day) iEEG monitoring is also interesting, but will require additional study. We postulate that in the future, combining pathological event classification, behavioral state classification, and multiple imaging modalities with interictal electrophysiological biomarkers could further improve interictal localization potential. We also believe that the insights drawn from this AI-based study will be particularly useful for clinical problems that require real-time data collection and decision making. References Amiri, M., Frauscher, B., & Gotman, J. Phase-amplitude coupling is elevated in deep sleep and in the in Human Neuroscience, 10, 387. onset http://doi.org/10.3389/fnhum.2016.00387 (2016). seizures. Frontiers zone of focal epileptic This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Ashburner, J., Barnes, G., Chen, C., Daunizeau, J., Flandin, G., Friston, K., Gitelman, D., Kiebel, S., Kilner, J., Litvak, V. and Moran, R., 2008. SPM8 manual. Functional Imaging Laboratory, Institute of Neurology, 41. Balach, J., et al. Comparison of algorithms for detection of high frequency oscillations in intracranial EEG. Proc. 2014 IEEE International Symposium on Medical Measurements and Applications (MeMeA) 1 -- 4. http://doi.org/10.1109/MeMeA.2014.6860107 (2014). Barkmeier, D. T., et al. High inter-reviewer variability of spike detection on intracranial EEG addressed by an automated multi-channel algorithm. Clinical Neurophysiology, 123(6), 1088 -- 1095. http://doi.org/10.1016/j.clinph.2011.09.023 (2012). Ben-Ari, Y., & Bragin, A. Physiologic and pathologic oscillations. Trends in Neurosciences, 30(7), 307 -- 308. http://doi.org/10.1016/j.tins.2007.05.008 (2007). 9(2), 137 -- 142. Blakely, T., Miller, K. J., Rao, R. P. N., Holmes, M. D., & Ojemann, J. G. Localization and classification of phonemes using high spatial resolution electrocorticography (ECoG) grids. Proc. 2008 30th Annual International Conference of the IEEE Engineering in Medicine and Biology Society, 4964 -- 4967. http://doi.org/10.1109/IEMBS.2008.4650328 (2008). Boser, B. E., Guyon, I. M. & Vapnik, V. N. A training algorithm for optimal margin classifiers. Proc. Fifth Annual Workshop on Computational Learning Theory, 144 -- 152. (1992). Bragin, A., Engel, J., Wilson, C. L., Fried, I., & Buzski, G. High-frequency oscillations in human brain. Hippocampus, http://doi.org/10.1002/(SICI)1098-1063(1999)9:2<137::AID- HIPO5>3.0.CO;2-0 (1999). Bragin, A., Wilson, C. L., Staba, R. J., Reddick, M., Fried, I., & Engel, J. Interictal high-frequency oscillations (80-500 Hz) in the human epileptic brain: Entorhinal cortex. Ann Neurol, 52, 407 -- 415. (2002). Brinkmann, B. H., Bower, M. R., Stengel, K. A., Worrell, G. A., & Stead, M. Large-scale electrophysiology: Acquisition, compression, encryption, and storage of big data. Journal of Neuroscience Methods, 180(1), 185 -- 192. http://doi.org/10.1016/j.jneumeth.2009.03.022 (2009). Brodie, M. J., & Kwan, P. Staged approach to epilepsy management. Neurology, 58(8 Suppl 5), S2 -- S8. http://doi.org/10.1212/WNL.58.8_SUPPL_5.S2 (2002). Burkholder, D. B., Sulc, V., Hoffman, E. M., Cascino, G. D., Britton, J. W., So, E. L., Marsh, W. R., Meyer, F. B., Van Gompel, J. J., Giannini, C., Wass, C. T., Watson, R. E., & Worrell G. A. Interictal scalp electroencephalography and intraoperative electrocorticography in magnetic resonance imaging-negative temporal lobe epilepsy surgery. JAMA Neurology, 71(6), 702 -- 709. (2014). Burnos, S., et al. Human intracranial high frequency oscillations (HFOs) detected by automatic time- frequency analysis. PLoS ONE, 9(4), e94381. http://doi.org/10.1371/journal.pone.0094381 (2014). Canolty, R. T., Cadieu, C. F., Koepsell, K., Knight, R. T., & Carmena, J. M. Multivariate phase -- amplitude cross-frequency coupling in neurophysiological signals. IEEE Transactions on Biomedical Engineering, 59(1), 8 -- 11. http://doi.org/10.1109/TBME.2011.2172439 (2012). Canolty, R. T., et al. High gamma power is phase-locked to theta oscillations in human neocortex. Science, 313(5793), 1626 -- 1628. http://doi.org/10.1126/science.1128115 (2006). Cherkassky, V. and Dhar, S., 2010, July. Simple Method for Interpretation of High-Dimensional Nonlinear SVM Classification Models. In DMIN (pp. 267-272). Cimbalnik, J., Kucewicz, M. T., & Worrell, G. Interictal high-frequency oscillations in focal human epilepsy. 175 -- 181. http://doi.org/10.1097/WCO.0000000000000302 (2016). Neurology, Opinion Current in 29(2), This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 & 80(7), of Neurology, Psychiatry, Neurosurgery Colgin, L. L. Rhythms of the hippocampal network. Nature Reviews Neuroscience, 17(4), 239 -- 249. http://doi.org/10.1038/nrn.2016.21 (2016). Cortes, C. & Vapnik, V. Support-vector networks. Machine Learning, 20(3), 273 -- 297. (1995). Crisler, S., Morrissey, M. J., Anch, A. M., & Barnett, D. W. Sleep-stage scoring in the rat using a support vector machine. Journal of Neuroscience Methods, 168(2), 524 -- 534. (2008). D'Alessandro, M., et al. Epileptic seizure prediction using hybrid feature selection over multiple intracranial EEG electrode contacts: A report of four patients. IEEE Transactions on Biomedical Engineering, 50(5), 603 -- 615. http://doi.org/10.1109/TBME.2003.810706 (2003). DiLorenzo, D. J., Mangubat, E. Z., Rossi, M. A., & Byrne, R. W. Chronic unlimited recording electrocorticography-guided resective epilepsy surgery: Technology-enabled enhanced fidelity in seizure focus localization with improved surgical efficacy. Journal of Neurosurgery, 120(6), 1402 -- 1414. http://doi.org/10.3171/2014.1.JNS131592 (2014). Edakawa, K., et al. Detection of epileptic seizures using phase-amplitude coupling in intracranial electroencephalography. Scientific Reports, 6(1), 25422. http://doi.org/10.1038/srep25422 (2016). Elsharkawy, A. E., et al. Long-term outcome of lesional posterior cortical epilepsy surgery in adults. Journal 773 -- 780. http://doi.org/10.1136/jnnp.2008.164145 (2009). Engel Jr, J., Bragin, A., Staba, R., & Mody, I. High-frequency oscillations: What is normal and what is not? Epilepsia, 50(4), 598 -- 604. http://doi.org/10.1111/j.1528-1167.2008.01917.x (2009). Engel, J., et al. Epilepsy biomarkers. Epilepsia, 54(Suppl. 4), 61 -- 69. http://doi.org/10.1111/epi.12299 (2013). Esteva, A., Kuprel, B., Novoa, R. A., Ko, J., Swetter, S. M., Blau, H. M., & Thrun, S. Dermatologist-level classification of skin cancer with deep neural networks. Nature, 542(7639), 115 -- 118. (2017). Fisher, R. S. & Velasco, A. L. Electrical brain stimulation for epilepsy. Nat Rev Neurol, 10, 261 -- 270. (2014). Gnatkovsky, V., de Curtis, M., Pastori, C., Cardinale, F., Lo Russo, G., Mai, R., Nobili, L., Sartori, I., Tassi, L., & Francione, S. Biomarkers of epileptogenic zone defined by quantified stereo-EEG analysis. Epilepsia, 55(2), 296 -- 305. (2014). Graef, A., et al. Automatic ictal HFO detection for determination of initial seizure spread. Proc. 2013 35th Annual International Conference of the IEEE Engineering in Medicine and Biology Society, 2096 -- 2099. https://doi.org/10.1109/EMBC.2013.6609946 (2013). Gritsch, G., et al. Automatic detection of the seizure onset zone based on ictal EEG. Proc. 2011 Annual International Conference of the IEEE Engineering in Medicine and Biology Society, 3901 -- 3904. http://doi.org/10.1109/IEMBS.2011.6090969 (2011). Gump, W. C., Skjei, K. L., & Karkare, S. N. Seizure control after subtotal lesional resection. Neurosurgical Focus, 34(6), E1. http://doi.org/10.3171/2013.3.FOCUS1348 (2013). Gulshan, V., Peng, L., Coram, M., Stumpe, M. C., Wu, D., Narayanaswamy, A., Venugopalan, S., Widner, K., Madams, T., Cuadros, J., & Kim, R. Development and validation of a deep learning algorithm for detection of diabetic retinopathy in retinal fundus photographs. JAMA, 316(22), 2402 -- 2410. (2016). Hauf, M., et al. Localizing seizure-onset zones in presurgical evaluation of drug-resistant epilepsy by electroencephalography/fMRI: Effectiveness of alternative thresholding strategies. American Journal of Neuroradiology, 33(9), 1818 -- 1824. http://doi.org/10.3174/ajnr.A3052 (2012). This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Höller, Y., et al. High-frequency oscillations in epilepsy and surgical outcome: A meta-analysis. Frontiers in Human Neuroscience, 9, 574. http://doi.org/10.3389/fnhum.2015.00574 (2015). Jacobs, J., LeVan, P., Chander, R., Hall, J., Dubeau, F., & Gotman, J. Interictal high-frequency oscillations (80-500 Hz) are an indicator of seizure onset areas independent of spikes in the human epileptic brain. Epilepsia, 49(11), 1893 -- 1907. http://doi.org/10.1111/j.1528-1167.2008.01656.x (2008). Jacobs, J., et al. High-frequency electroencephalographic oscillations correlate with outcome of epilepsy surgery. Annals of Neurology, 67(2), 209 -- 220. http://doi.org/10.1002/ana.21847 (2010). Jenks, G. F. The data model concept in statistical mapping. International Yearbook of Cartography, 7, 186 -- 190. (1967). Jensen, O., & Colgin, L. L. Cross-frequency coupling between neuronal oscillations. Trends in Cognitive Sciences, 11(7), 267 -- 269. http://doi.org/10.1016/j.tics.2007.05.003 (2007). Jirsch, J. D., et al. High-frequency oscillations during human focal seizures. Brain, 129(6), 1593 -- 1608. http://doi.org/10.1093/brain/awl085 (2006). Karoly, P. J., et al. Interictal spikes and epileptic seizures: Their relationship and underlying rhythmicity. Brain, 139(4), 1066 -- 1078. http://doi.org/10.1093/brain/aww019 (2016). Kremen, V., et al. Behavioral state classification in epileptic brain using intracranial electrophysiology. Journal of Neural Engineering, 14(2), 26001. http://doi.org/10.1088/1741-2552/aa5688 (2017). Kucewicz, M. T., et al. Dissecting gamma frequency activity during human memory processing. Brain, 140(5), 1337 -- 1350. http://doi.org/10.1093/brain/awx043 (2017). Kwan, P., & Brodie, M. J. Early identification of refractory epilepsy. The New England Journal of Medicine, 342(5), 314 -- 319. http://doi.org/10.1056/NEJM200002033420503 (2000). Lee, S. K. Surgical approaches in nonlesional neocortical epilepsy. Journal of Epilepsy Research, 1(2), 47 -- 51. http://doi.org/10.14581/jer.11009 (2011). Leonardi, M., & Ustun, T. B. The global burden of epilepsy. Epilepsia, 43(s6), 21 -- 25. (2002). Liu, S., et al. Exploring the time -- frequency content of high frequency oscillations for automated identification of seizure onset zone in epilepsy. Journal of Neural Engineering, 13(2), 26026. http://doi.org/10.1088/1741-2560/13/2/026026 (2016). Lüders, H. O., Najm, I., Nair, D., Widdess-Walsh, P., & Bingman, W. The epileptogenic zone: General principles. Epileptic Disorders, http://www.jle.com/en/revues/epd/e- docs/the_epileptogenic_zone_general_principles_270796/article.phtml (2006). 8(Suppl. 2), S1 -- S9. Luther, N., Rubens, E., Sethi, N., Kandula, P., Labar, D. R., Harden, C., Perrine, K., Christos, P. J., Iorgulescu, J. B., Lancman, G., Schaul, N. S., Kolesnik, D. V., Nouri, S., Dawson, A., Tsiouris, A. J., & Schwartz, T. H. The value of intraoperative electrocorticography in surgical decision making for temporal lobe epilepsy with normal MRI. Epilepsia, 52(5), 941 -- 948. (2011). Malow, B. A., Lin, X., Kushwaha, R., & Aldrich, M. S. Interictal spiking increases with sleep depth in temporal lobe epilepsy. Epilepsia, 39(12), 1309 -- 1316. http://doi.org/10.1111/j.1528-1157.1998.tb01329.x (1998). Marsh, E. D., et al. Interictal EEG spikes identify the region of electrographic seizure onset in some, but not all, pediatric epilepsy patients. Epilepsia, 51(4), 592 -- 601. http://doi.org/10.1111/j.1528- 1167.2009.02306.x (2010). Matsumoto, A., et al. Pathological and physiological high-frequency oscillations in focal human epilepsy. Journal of Neurophysiology, 110(8), 1958 -- 1964. http://doi.org/10.1152/jn.00341.2013 (2013). This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Miller, K. J., Zanos, S., Fetz, E. E., den Nijs, M., & Ojemann, J. G. Decoupling the cortical power spectrum reveals real-time representation of individual finger movements in humans. Journal of Neuroscience, 29(10), 3132 -- 3137. http://doi.org/10.1523/JNEUROSCI.5506-08.2009 (2009). Nair, C., Prabhakar, B., & Shah, D. On entropy for mixtures of discrete and continuous variables. Retrieved from http://arxiv.org/abs/cs/0607075 (2006). Noe, K., et al. Long-term outcomes after nonlesional extratemporal lobe epilepsy surgery. JAMA Neurology, 70(8), 1003 -- 1008. http://doi.org/10.1001/jamaneurol.2013.209 (2013). Nonoda, Y., et al. Interictal high-frequency oscillations generated by seizure onset and eloquent areas may be differentially coupled with different slow waves. Clinical Neurophysiology, 127(6), 2489 -- 2499. http://doi.org/10.1016/j.clinph.2016.03.022 (2016). Pail, M., Řehulka, P., Cimbálník, J., Doležalová, I., Chrastina, J., & Brázdil, M. Frequency-independent in epileptic and non-epileptic regions. Clinical characteristics of high-frequency oscillations Neurophysiology, 128(1), 106 -- 114. http://doi.org/10.1016/j.clinph.2016.10.011 (2017). Papadelis, C., simultaneous magnetoencephalography and electroencephalography as biomarker of pediatric epilepsy. Journal of Visualized Experiments, 118, e54883. http://doi.org/10.3791/54883 (2016). Pearce, A., Wulsin, D., Blanco, J. A., Krieger, A., Litt, B., & Stacey, W. C. Temporal changes of neocortical high frequency oscillations in epilepsy. J Neurophysiol, 110(5), 1167 -- 1179. (2013). Refaeilzadeh, P., Tang, L., & Liu, H. Cross-validation. In Encyclopedia of Database Systems, L. Liu & M. T. Özsu, Eds. (pp. 532 -- 538). Springer US. (2009). Russell, S. J. and Norvig, P. Artificial Intelligence: A Modern Approach. Prentice-Hall, Englewood Cliffs. (1995). Sammaritano, M., Gigli, G. L., & Gotman, J. Interictal spiking during wakefulness and sleep and the localization of foci in temporal lobe epilepsy. Neurology, 41(2(Pt 1)), 290 -- 297. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/1992379 (1991). frequency oscillations detected with et al. Interictal high Schwartz, T.H., Bazil, C.W., Walczak, T.S., Chan, S., Pedley, T.A., and Goodman, R.R. (1997). The predictive value of intraoperative electrocorticography in resections for limbic epilepsy associated with mesial temporal sclerosis. Neurosurgery 40, 302-9; discussion 309-11. Sinha, N., et al. Predicting neurosurgical outcomes in focal epilepsy patients using computational modelling. Brain, 140(2), 319 -- 332. http://doi.org/10.1093/brain/aww299 (2017). Staba, R. J., Wilson, C. L., Bragin, A., Fried, I., & Engel, J. Quantitative analysis of high-frequency oscillations (80-500 Hz) recorded in human epileptic hippocampus and entorhinal cortex. Journal of Neurophysiology, 88(4), 1743 -- 1752. https://doi.org/10.1152/jn.2002.88.4.1743 (2002). Staba, R. J., Wilson, C. L., Bragin, A., Jhung, D., Fried, I., & Engel, J. High-frequency oscillations recorded in human medial temporal lobe during sleep. Annals of Neurology, 56(1), 108 -- 115. http://doi.org/10.1002/ana.20164 (2004). Staley, K. J., & Dudek, F. E. Interictal spikes and epileptogenesis. Epilepsy Currents, 6(6), 199 -- 202. http://doi.org/10.1111/j.1535-7511.2006.00145.x (2006). Stead, M., Bower, M., Brinkmann, B. H., Lee, K., Marsh, W. R., Meyer, F. B., Litt, B., Van Gompel, J., & Worrell, G. A. Microseizures and the spatiotemporal scales of human partial epilepsy. Brain, 133(9), 2789 -- 2797. (2010). Tellez-Zenteno, J. F. Long-term seizure outcomes following epilepsy surgery: A systematic review and meta-analysis. Brain, 128(5), 1188 -- 1198. http://doi.org/10.1093/brain/awh449 (2005). This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Tzallas, A. T., Tsipouras, M. G., & Fotiadis, D. I., Automatic seizure detection based on time-frequency analysis and artificial neural networks. Computational Intelligence and Neuroscience, 2007, 80510. https://www.hindawi.com/journals/cin/2007/080510/abs/ (2007). Van Gompel, J. J., Worrell, G. A., Bell, M. L., Patrick, T. A., Cascino, G. D., Raffel, C., Marsh, W. R., & Meyer, F. B. Intracranial electroencephalography with subdural grid electrodes: Techniques, complications, and outcomes. Neurosurgery, 63(3), 498 -- 505; discussion 505 -- 506. (2008). Van Gompel, J. J., Rubio, J., Cascino, G. D., Worrell, G. A., & Meyer, F. B. Electrocorticography-guided resection of temporal cavernoma: Is electrocorticography warranted and does it alter the surgical approach? J Neurosurg, 110(6), 1179 -- 1185. (2009). van 't Klooster, M. A., Leijten, F. S., Huiskamp, G., Ronner, H. E., Baayen, J. C., van Rijen, P. C., Eijkemans, M. J., Braun, K. P., Zijlmans, M., & HFO study group. High frequency oscillations in the intra-operative ECoG to guide epilepsy surgery ("The HFO Trial"): Study protocol for a randomized controlled trial. Trials, 16, 422. (2015). Vapnik, V. Estimation of Dependences Based on Empirical Data [2nd Ed.]: Empirical Inference Science Afterword of 2006. New York, NY: Springer. https://doi.org/10.1007/0-387-34239-7 (2006). Varatharajah, Y., Berry, B.M., Kalbarczyk, Z.T., Brinkmann, B.H., Worrell, G.A. and Iyer, R.K., 2017, May. Inter-ictal Seizure Onset Zone localization using unsupervised clustering and Bayesian Filtering. In Neural Engineering (NER), 2017 8th International IEEE/EMBS Conference on (pp. 533-539). IEEE. Varatharajah, Y., Chong, M. J., Saboo, K., Berry, B., Brinkmann, B., Worrell, G., & Iyer, R. EEG- GRAPH: A factor-graph-based model for capturing spatial, temporal, and observational relationships in electroencephalograms. Advances in Neural Information Processing Systems, 30, 5377 -- 5386. https://papers.nips.cc/book/advances-in-neural-information-processing-systems-30-2017 (2017). von Ellenrieder, N., Dubeau, F., Gotman, J., & Frauscher, B. Physiological and pathological high- frequency oscillations have distinct sleep-homeostatic properties. NeuroImage: Clinical, 14, 566 -- 573. http://doi.org/10.1016/j.nicl.2017.02.018 (2017). Wass, C.T., Grady, R.E., Fessler, A.J., Cascino, G.D., Lozada, L., Bechtle, P.S., Marsh, W.R., Sharbrough, F.W., and Schroeder, D.R. (2001). The effects of remifentanil on epileptiform discharges during intraoperative electrocorticography in patients undergoing epilepsy surgery. Epilepsia 42, 1340- 1344. Weiss, S. A., et al. Ictal high frequency oscillations distinguish two types of seizure territories in humans. Brain, 136(12), 3796 -- 3808. http://doi.org/10.1093/brain/awt276 (2013). Weiss, S. A., et al. Seizure localization using ictal phase-locked high gamma: A retrospective surgical outcome study. Neurology, 84(23), 2320 -- 2328. http://doi.org/10.1212/WNL.0000000000001656 (2015). Weiss, S.A., Alvarado‐Rojas, C., Bragin, A., Behnke, E., Fields, T., Fried, I., Engel, J., & Staba, R. Ictal onset patterns of local field potentials, high frequency oscillations, and unit activity in human mesial temporal lobe epilepsy. Epilepsia, 57(1), 111 -- 121. (2016). Wellmer, J., von der Groeben, F., Klarmann, U., Weber, C., Elger, C. E., Urbach, H., Clusmann, H., von Lehe, M. Risks and benefits of invasive epilepsy surgery workup with implanted subdural and depth electrodes. Epilepsia, 53, 1322 -- 1332. (2012). Worrell, G. A., et al. High-frequency oscillations and seizure generation in neocortical epilepsy. Brain, 127(7), 1496 -- 1506. http://doi.org/10.1093/brain/awh149 (2004). Worrell, G. A., et al. Recording and analysis techniques for high-frequency oscillations. Progress in Neurobiology, 98(3), 265 -- 278. http://doi.org/10.1016/j.pneurobio.2012.02.006 (2012). This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960 Worrell, G., & Gotman, J. High-frequency oscillations and other electrophysiological biomarkers of epilepsy: Clinical studies. Biomarkers in Medicine, 5(5), 557 -- 566. http://doi.org/10.2217/bmm.11.74 (2011). Wu, J. Y., et al. Removing interictal fast ripples on electrocorticography linked with seizure freedom in children. Neurology, 75(19), 1686 -- 1694. http://doi.org/10.1212/WNL.0b013e3181fc27d0 (2010). Ylinen, A., et al. Sharp wave-associated high-frequency oscillation (200 Hz) in the intact hippocampus: intracellular mechanisms. The Network 30 -- 46. http://www.jneurosci.org/content/15/1/30.long (1995). Zijlmans, M., et al. High-frequency oscillations as a new biomarker in epilepsy. Annals of Neurology, 71(2), 169 -- 178. http://doi.org/10.1002/ana.22548 (2012). of Neuroscience, and Journal 15(1), This manuscript has been published in the Journal of Neural Engineering. DOI: 10.1088/1741-2552/aac960
1501.00727
3
1501
2015-01-12T10:07:48
Graph Theoretical Analysis Reveals: Women's Brains are Better Connected than Men's
[ "q-bio.NC" ]
Deep graph-theoretic ideas in the context with the graph of the World Wide Web led to the definition of Google's PageRank and the subsequent rise of the most-popular search engine to date. Brain graphs, or connectomes, are being widely explored today. We believe that non-trivial graph theoretic concepts, similarly as it happened in the case of the World Wide web, will lead to discoveries enlightening the structural and also the functional details of the animal and human brains. When scientists examine large networks of tens or hundreds of millions of vertices, only fast algorithms can be applied because of the size constraints. In the case of diffusion MRI-based structural human brain imaging, the effective vertex number of the connectomes, or brain graphs derived from the data is on the scale of several hundred today. That size facilitates applying strict mathematical graph algorithms even for some hard-to-compute (or NP-hard) quantities like vertex cover or balanced minimum cut. In the present work we have examined brain graphs, computed from the data of the Human Connectome Project, recorded from male and female subjects between ages 22 and 35. Significant differences were found between the male and female structural brain graphs: we show that the average female connectome has more edges, is a better expander graph, has larger minimal bisection width, and has more spanning trees than the average male connectome. Since the average female brain weights less than the brain of males, these properties show that the female brain is more "well-connected" or perhaps, more "efficient" in a sense than the brain of males.
q-bio.NC
q-bio
Graph Theoretical Analysis Reveals: Women's Brains are Better Connected than Men's Bal´azs Szalkaia, B´alint Vargaa, Vince Grolmusza,b,∗ aPIT Bioinformatics Group, Eotvos University, H-1117 Budapest, Hungary bUratim Ltd., H-1118 Budapest, Hungary Abstract Deep graph-theoretic ideas in the context with the graph of the World Wide Web led to the definition of Google's PageRank and the subsequent rise of the most- popular search engine to date. Brain graphs, or connectomes, are being widely explored today. We believe that non-trivial graph theoretic concepts, similarly as it happened in the case of the World Wide web, will lead to discoveries enlightening the structural and also the functional details of the animal and human brains. When scientists examine large networks of tens or hundreds of millions of vertices, only fast algorithms can be applied because of the size constraints. In the case of diffusion MRI-based structural human brain imaging, the effective vertex number of the connectomes, or brain graphs derived from the data is on the scale of several hundred today. That size facilitates applying strict mathematical graph algorithms even for some hard-to-compute (or NP- hard) quantities like vertex cover or balanced minimum cut. In the present work we have examined brain graphs, computed from the data of the Human Connectome Project, recorded from male and female subjects between ages 22 and 35. Significant differences were found between the male and female structural brain graphs: we show that the average female connectome has more edges, is a better expander graph, has larger minimal bisection width, and has more spanning trees than the average male connectome. Since the average female brain weights less than the brain of males, these properties show that the female brain is more "well-connected" or perhaps, more "efficient" in a sense than the brain of males. It is known that the female brain has a larger white matter/gray matter ratio than the brain of males; this observation is in line with our findings concerning the number of edges, since the white matter consists of myelinated axons, which, in turn, correspond to the connections in the brain graph. We have also found that the minimum bisection width, normalized with the edge number, is also significantly larger in the right and the left hemispheres in females: therefore, that structural difference is independent from the difference in the number of edges. ∗Corresponding author Email addresses: [email protected] (Bal´azs Szalkai), [email protected] (B´alint Varga), [email protected] (Vince Grolmusz) 1 1. Introduction In the last several years hundreds of publications appeared describing or analyzing structural or functional networks of the brain, frequently referred to as "connectome" [1, 2]. Some of these publications analyzed data from healthy humans [3, 4, 5, 6], and some compared the connectome of the healthy brain with diseased one [7, 8, 9, 10, 11]. So far, the analyses of the connectomes mostly used tools developed for very large networks, such as the graph of the World Wide Web (with billions of vertices), or protein-protein interaction networks (with tens or hundreds of thou- sands of vertices), and because of the huge size of original networks, these meth- ods used only very fast algorithms and frequently just primary degree statistics and graph-edge counting between pre-defined regions or lobes of the brain [12]. In the present work we demonstrate that deep and more intricate graph the- oretic parameters could also be computed by using, among other tools, contem- porary integer programming approaches for connectomes with several hundred vertices. With these mathematical tools we show statistically significant differences in some graph properties of the connectomes, computed from MRI imaging data of male and female brains. We will not try to associate behavioral patterns of males and females with the discovered structural differences [12] (see also the debate that article has generated: [13, 14, 15]), because we do not have behavioral data of the subjects of the imaging study, and, additionally, we cannot describe high- level functional properties implied by those structural differences. However, we clearly demonstrate that deep graph-theoretic parameters show "better" connections in a certain sense in female connectomes than in male ones. The study of [12] analyzed the 95-vertex graphs of 949 subjects aged be- tween 8 and 22 years, using basic statistics for the numbers of edges running either between or within different lobes of the brain (the parameters deduced were called hemispheric connectivity ratio, modularity, transitivity and partic- ipation coefficients, see [12] for the definitions). It was found that males have significantly more intra-hemispheric edges than females, while females have sig- nificantly more inter-hemispheric edges than males. 2. Results and Discussion We have analyzed the connectomes of 96 subjects, 52 females and 44 males, each with 83, 129 and 234 node resolutions, and each graphs with five different weight functions. We considered the connectomes as graphs with weighted edges, and performed graph-theoretic analyses with computing some polynomial-time computable and also some NP-hard graph parameters on the individual graphs, and then compared the results statistically for the male and the female group. 2 We have found that female connectomes have more edges, larger (normal- ized) minimum bisection widths, larger minimum-vertex covers and more span- ning trees than the male connectomes. In order to describe the parameters, which differ significantly among male and female connectomes, we need to place them in the context of their graph theoretical definitions. 2.1. Edge number and edge weights We have found significantly higher number of edges (counted with 5 types of weights and also without any weights) in both hemispheres and also in the whole brain in females, in all resolutions. This finding is surprising, since we used the same parcellation and the same tractography and the same graph- construction methods for female and male brains, and because it is proven that females have, on average, less-weighting brains than males [16]. For example, in the 234-vertex resolution, the average number of (unweighted) edges in female connectomes is 1826, in males 1742, with p = 0.00063 (see the Appendix for tables with the results). The work of [12] reported similar findings in inter- hemispheric connections only. It is known that there are statistical differences in the size and the weight of the female and the male cerebra [16]. It was also published [17] that female brains statistically have a higher white matter/gray matter ratio than male brains. We argue that this observation is in line with the quantitative differences in the fibers and edges in the connectomes of the sexes. In a simplified view, the edges of the braingraph correspond to the fibers of the myelinated axons in the white matter, while the nodes of the graph to areas of the gray matter. Therefore, since females have a higher white matter/gray matter ratio than males by [17] that fact implies that the number of detected fibers by the tractography step of the processing is relatively higher in females than in males, and this higher number of fibers imply higher number of edges in female connectomes. 2.2. Minimum cut and balanced minimum cut Suppose the nodes, or the vertices, of a graph are partitioned into two, disjoint, non-empty sets, say X and Y ; their union is the whole vertex-set of the graph. The X, Y cut is the set of all edges connecting vertices of X with the vertices of Y (Figure 1A). The size of the cut is the number of edges in the cut. In graph theory, the size of the minimum cut is an interesting quantity. The minimum cut between vertices a and b is the minimum cut, taken for all X and Y , where vertex a is in X and b is in Y . This quantity gives the "bottleneck", in a sense, between those two nodes (c.f., Menger theorems and Ford-Fulkerson's Min-Cut-Max-Flow theorem [18, 19]). The minimum cut in a graph is defined to be the cut with the fewest edges for all non-empty sets X and Y , partitioning the vertices. Clearly, for non-negative weights, the size of the minimum cut in a non- connected graph is 0. Very frequently, however, in connected graphs, the mini- mum cut is determined by just the smallest degree node: that node is the only 3 element of set X and all the other vertices of the graph are in Y (Figure 1B). Because of this phenomenon, the minimum cut is frequently queried for the "balanced" case, when the size (i.e., the number of vertices) of X and Y needs to be equal (or, more exactly, may differ by at most one if the number of the vertices of the graph is odd), see Figure 1C. This problem is referred to as the balanced minimum cut or the minimum bisection problem. If the minimum bi- section is small that means that there exist a partition of the vertices into two sets of equal size that are connected with only a few edges. If the minimum bisection is large then the two half-sets in every possible bisections of the graph are connected by many edges. Therefore, the balanced minimum cut of a graph is independent of the par- ticular labeling of the nodes. The number of all the balanced cuts in a graph with n vertices is greater than 1 n + 1 2n, that is, for n = 250, this number is very close to the number of atoms in the visible universe [20]. Consequently, one cannot practically compute the minimum bisecton width by reviewing all the bisectons in a graph of that size. Moreover, the complexity of computing this quantity is known to be NP-hard [21] in general, but with contemporary integral programming approaches, for the graph-sizes we are dealing with, the exact values are computable. In computer engineering, an important measure of the quality of an inter- connection network is its minimum bisection width [22]: the higher the width is the better the network. For the whole brain graph, as it is anticipated, we have found that the minimum balanced cut is almost exactly represents the edges crossing the corpus callosum, connecting the two cerebral hemispheres. We show that within both hemispheres, the minimum bisection size of female connectomes are significantly larger than the minimum bisection size of the males. Much more importantly, we show that this remains true if we normalize with the sum of all edge-weights: that is, this phenomenon cannot be due to the higher number of edges or the greater edge weights in the female brain: it is an intrinsic property of the female brain graph in our data analyzed. For example, in the 234-vertex resolution, in the left hemisphere, the nor- malized balanced minimum cut in females, on the average, is 0.09416, in the males 0.07896, p = 0.00153 (see the Appendix for tables with the results). We think that this finding is one of the main results of the present work: even if the significant difference in the weighted edge numbers are due to some artifacts in the data acquisition/processing workflow, the normalized balanced minimum cut size seems to be independent from those processes. 2.3. Eigengap and the expander property Expander graphs and the expander-property of graphs are one of the most interesting area of graph theory: they are closely related to the convergence rate and the ergodicity of Markov chains, and have applications in the design 4 Figure 1: Panel A: An X-Y cut. The cut-edges are colored black. Panel B: An un-balanced minimum cut. Panel C: A balanced cut. Panel D: The wheel graph. of communication- and sorting networks and methods for de-randomizing algo- rithms [23]. A graph is an ε-expander, if every – not too small and not too large – vertex-set S of the graph has at least εS outgoing edges (see [23] for the exact definition). Random walks on good expander graphs converge very fast to the limit distribution: this means that good expander graphs, in a certain sense, are "intrinsically better" connected than bad expanders. It is known that large eigengap of the walk transition matrix of the graph implies good expansion property [23]. We have found that women's connectomes have significantly larger eigengap, and, consequently, they are better expander graphs than the connectomes of men. For example, in the 83-node resolution, in the left hemisphere and in the unweighted graph, the average female connectome's eigengap is 0.306 while in the case of men it is 0.272, with p = 0.00458. 2.4. The number of spanning forests A tree in graph theory is a connected, cycle-free graph. Any tree on n vertices has the same number of edges: n − 1. Trees, and tree-based structures are common in science: phylogenetic trees, hierarchical clusters, data-storage on hard-disks, or a computational model called decision trees all apply graph- theoretic trees. A spanning tree is a minimal subgraph of a connected graph that is still connected. Some graphs have no spanning trees at all: only connected 5 graphs have spanning trees. A tree has only one spanning tree: itself. Any connected graph on n vertices has a minimum of n− 1 and a maximum of n(n− 1)/2 edges [24]. A connected graph with few edges still may have exponentially many different spanning trees: e.g., the n-vertex wheel on Figure 1D has at least 2n−1 spanning trees (for n ≥ 4). Cayley's famous theorem, and its celebrated proof with Prufer codes [25] shows that the number of spanning trees of the complete graph on n vertices is nn−2. If a graph is not connected, then it contains more than one connected com- ponents. Each connected component has at least one spanning tree, and the whole graph has at least one spanning forest, comprising of the spanning trees of the components. The number of spanning forests is clearly the product of the numbers of the spanning trees of the components. For graphs in general, one can compute the number of their spanning forests by Kirchoff's matrix tree theorem [26, 27] using the eigenvalues of the Laplacian matrix [27] of the graph. We show that female connectomes have significantly higher number of span- ning trees than the connectomes of males. For example, in the 129-vertex resolu- tion, in the left hemisphere, the logarithm of the number of the spanning forests in the unweighted case are 162.01 in females, 158.88 in males with p = 0.013. 3. Materials and Methods 3.1. Data source and graph computation: The dataset applied is a subset of the Human Connectome Project [28] anonymized 500 Subjects Release: (http://www.humanconnectome.org/documentation/S500) of healthy subjects between 22 and 35 years of age. Data was downloaded in October, 2014. The Connectome Mapper Toolkit [29] (http://cmtk.org) was applied for brain tissue segmentation into grey and white matter, partitioning, tractography and the construction of the graphs from the fibers identified in the tractography step. The Connectome Mapper Toolkit [29] default partitioning was used (computed by the FreeSurfer, and based on the Desikan-Killiany anatomical atlas) into 83, 129 and 234 cortical and sub-cortical structures (as the brainstem and deep- grey nuclei), referred to as "Regions of Interest", ROIs, (see Figure 4 in [29]). Tractography was performed by choosing the deterministic streamline method [29] with randomized seeding. The graphs were constructed as follows: the nodes correspond to the ROIs in the specific resolution. Two nodes were connected by an edge if there exists at least one fiber (determined by the tractography step) connecting the ROIs, corresponding to the nodes. More than one fibers, connecting the same nodes, may give rise to the weight of that edge, depending on the weighting method. Loops were deleted from the graph. The weights of the edges are assigned by several methods, taking into account the lengths and the multiplicities of the fibers, connecting the nodes: • Unweighted: Each edge has weight 1. 6 • FiberN: The number of fibers traced along the edge: this number is larger than one if more than one fibers connect two cortical or sub-cortical areas, corresponding to the two endpoints of the edge. • FAMean: The arithmetic mean of the fractional anisotropies [30] of the fibers, belonging to the edge. • FiberLengthMean: The average length of the fibers, connecting the two endpoints of the edge. • FiberNDivLength: The number of fibers belonging to the edge, divided by their average length. This quantity is related to the simple electrical model of the nerve fibers: by modeling the fibers as electrical resistors with resistances proportional to the average fiber length, this quantity is precisely the conductance between the two regions of interest. Addi- tionally, FiberNDivLength can be observed as a reliability measure of the edge: longer fibers are less reliable than the shorter ones, due to possi- ble error accumulation in the tractography algorithm that constructs the fibers from the anisotropy data. Multiple fibers connecting the same two ROIs, corresponding to the endpoints, add to the reliability of the edge, because of the independently tractographed connections. By generalized adjacency matrix we mean a matrix of size n × n where n is the number of nodes (or vertices) in the graph, whose rows and columns correspond to the nodes, and whose each element is either zero if there is no edge between the two nodes, or equals to the weight of the edge connecting the two nodes. By the generalized degree of a node we mean the sum of the weights of the edges adjacent to that node. Note that the generalized degree of the node v is exactly the sum of the elements in the row (or column) of the generalized adjacency matrix corresponding to v. By generalized Laplacian matrix we mean the matrix D − A, where D is a diagonal matrix containing the generalized degrees, and A is the generalized adjacency matrix. 3.2. Graph parameters: We calculated various graph parameters for each brain graph and weight function. These parameters included: • Number of edges (Sum). The weighted version of this quantity is the sum of the weights of the edges. • Normalized largest eigenvalue (AdjLMaxDivD): The largest eigenvalue of the generalized adjacency matrix, divided by the average degree. Divid- ing by the average degree of vertices was necessary because the largest eigenvalue is bounded by the average- and maximum degrees, and thus is considered by some a kind of "average degree" itself [24]. This means that a denser graph may have a bigger λmax largest eigenvalue solely because of a larger average degree. We note that the average degree is already defined by the sum of weights. 7 • Eigengap of the transition matrix (PGEigengap): The transition matrix PG is obtained by dividing all the rows of the generalized adjacency matrix by the generalized degree of the corresponding node. When performing a random walk on the graph, for nodes i and j, the corresponding matrix element describes the probability of transitioning to node j, supposing that we are at node i. The eigengap of a matrix is the difference of the largest and the second largest eigenvalue. It is characteristic to the expander properties of the graph: the larger the gap, the better expander is the graph (see [23] for the exact statements and proofs). • Hoffman's bound (HoffmanBound): The expression 1 + λmax λmin , where λmax and λmin denote the largest and smallest eigenvalues of the adjacency matrix. It is a lower bound for the chromatic number of the graph. The chromatic number is generally higher for denser graphs, as the addition of an edge may make a previously valid coloring invalid. • Logarithm of number of spanning forests (LogAbsSpanningForestN): The number of the spanning trees in a connected graph can be calculated from the spectrum of its Laplacian [26, 27]. Denser graphs tend to have more spanning trees, as the addition of an edge introduces zero or more new spanning trees. If a graph is not connected, then the number of spanning forests is the product of the numbers of the spanning trees of the components. The parameter LogAbsSpanningForestN equals to the logarithm of the number of spanning forests in the unweighted case. In the case of other weight functions, if we define the weight of a tree by the product of the weights of its edges, then this parameter equals to the sum of the logarithms of the weights of the spanning trees in the forests. • Balanced divided by the cut, minimum of edges (MinCutBalDivSum): The task is to partition the graph into two sets whose size may differ from each other by at most 1, so that the num- ber of edges crossing the cut is minimal. This is the "balanced minimum cut" problem, or sometimes called the "minimum bisection width" prob- lem. For the whole brain graph, our expectation was that the minimum cut corresponds to the boundary of the two hemispheres, which was indeed proven when we analyzed the results. number • Minimum cost spanning tree (MinSpanningForest), calculated with Kruskal's algorithm. • Minimum weighted vertex cover (MinVertexCover): Each vertex should have a (possibly fractional) weight assigned such that, for each edge, the sum of the weights of its two endpoints is at least 1. This is the fractional relaxation of the NP-hard vertex-cover problem [31]. The minimum of the sum of all vertex-weights is computable by a linear programming approach. 8 • Minimum vertex cover (MinVertexCoverBinary): Same as above, but each weight must be 0 or 1. In other words, a minimum size set of vertices is selected such that each edge is covered by at least one of the selected ver- tices. This NP-hard graph-parameter is computed only for the unweighted case. The exact values are computed by an integer programming solver SCIP (http://scip.zib.de), [32, 33]. The above 9 parameters were computed for all three resolutions and for the left and the right hemispheres and also for the whole connectome, with all 5 weight functions (with the following exceptions: MinVertexCoverBinary was computed only for the unweighted case, and the MinSpanningForest was not computed for the unweighted case). 3.3. Statistical analysis Since each connectome was computed in multiple resolutions (in 83, 129 and 234 nodes), we had three graphs for each brain. In addition, the parameters were calculated separately for the connectome within the left and right hemi- spheres as well, not only the whole graph, since we intended to examine whether statistically significant differences can be attributed to the left or right hemi- spheres. Each subjects' brain was corresponded to 9 graphs (3 resolutions, each in the left and the right hemispheres, plus the whole cortex with sub-cortical areas) and for each graph we calculated 9 parameters, each (with the excep- tions noted above) with 5 different edge weights. This means that we assigned 7 · 5 · 3 + 1 · 3 + 4 · 3 = 120 attributes to each resolution of the 96 brains, that is, 360 attributes to each brain. The statistical null hypothesis [34] of ours was that the graph parameters do not differ between the male and the female groups. As the first approach, we have used ANOVA (Analysis of variance) [35] to assign p-values for all parame- ters in each hemispheres and in each resolutions and in each weight-assignments. Our very large number of attributes may lead to false negatives, i.e., to "type II" statistical errors: in other words, it may happen that an attribute, with a very small p-value may appear "at random", simply because we tested a lot of attributes. In order to deal with "type II" statistical errors, we followed the route described below. We divided the population randomly into two sets by the parity of the sum of the digits in their ID. The first set was used for making hypotheses and the second set for testing these hypotheses. This was necessary to avoid type II errors resulting from multiple testing correction. If we made hypotheses for all the numerical parameters, then the Holm-Bonferroni correction [36] we used would have unnecessarily increased the p-values. Thus we needed to filter the hypotheses first, and that is why we needed the first set. Testing on the first set allowed us to reduce the number of hypotheses and test only a few of them on the second set. The hypotheses were filtered by performing ANOVA (Analysis of variance) [35] on the first set. Only those hypotheses were selected to qualify for the second round where the p-value was less than 1%. The selected hypotheses 9 were then tested for the second set as well, and the resulting p-value corrected with the Holm-Bonferroni correction method [36] with a significance level of 5%. In Table 1 those hypotheses rejected were highlighted in bold, meaning that all the corresponding graph parameters differ significantly in sex groups at a combined significance level of 5%. We also highlighted (in italic) those p-values which were individually less than the threshold, meaning that these hypotheses can individually be rejected at a level of 5%, but it is very likely that not all of these graph parameters are significantly different between the sexes. 4. Conclusions: We have computed 83-, 129- and 234-vertex-graphs from the diffusion MRI images of the 96 subjects of 52 females and 44 males, between the age of 22 and 35. We have found, after a careful statistical analysis, significant differ- ences between some graph theoretical parameters of the male and female brain graphs. Our findings show that the female brain graphs have generally more edges (counted with and without weights), have larger normalized minimum bi- section widths and have more spanning trees (counted with and without weights) than the connectomes of males (Table 1). Additionally, with weaker statistical validity, some spectral properties and the minimum vertex cover also differ in the connectomes of different sexes (each with p < 0.02). 5. Data availability: The unprocessed and pre-processed MRI data is available at the Human Connectome Project's website: http://www.humanconnectome.org/documentation/S500 [28]. 5.1. Table 1 Scale 129 83 234 129 83 83 129 234 83 129 83 83 234 234 129 Property Right MinCutBalDivSum FAMean All LogSpanningForestN FiberNDivLength All PGEigengap FiberNDivLength All PGEigengap FiberNDivLength Left MinCutBalDivSum FiberN Right MinCutBalDivSum FAMean Left PGEigengap FiberNDivLength All PGEigengap FiberN All Sum Unweighted Left MinCutBalDivSum FiberN All LogSpanningForestN FiberN Right Sum FAMean All Sum Unweighted Left PGEigengap FiberNDivLength All Sum Unweighted 10 p (1st) 0.00807 0.00003 0.00321 0.00792 0.00403 0.00496 0.00223 0.00826 0.00025 0.00001 0.00001 0.00028 0.00063 0.00013 0.00026 p (2nd) 0.00003 0.00004 0.00007 0.00011 0.00011 0.00015 0.00015 0.00022 0.00022 0.00023 0.00028 0.00029 0.00032 0.00038 0.00042 p (corrected) 0.00401 0.00451 0.00798 0.01303 0.01300 0.01744 0.01797 0.02517 0.02504 0.02563 0.03084 0.03224 0.03512 0.04171 0.04563 234 129 83 129 234 83 234 234 129 234 83 83 234 83 83 129 83 129 234 234 129 83 129 129 129 129 129 234 83 234 83 129 83 234 234 234 234 83 83 129 83 234 83 234 234 129 83 234 234 129 All Sum FAMean All LogSpanningForestN FiberN All Sum FAMean Right Sum FAMean Right PGEigengap FiberNDivLength Left Sum Unweighted Right Sum FAMean Left Sum Unweighted All Sum FAMean Left MinCutBalDivSum FiberN Left LogSpanningForestN FiberNDivLength All LogSpanningForestN Unweighted Left MinCutBalDivSum FiberLengthMean All LogSpanningForestN FAMean Right Sum Unweighted Left MinCutBalDivSum Unweighted Left MinCutBalDivSum Unweighted Left PGEigengap FiberN All LogSpanningForestN FAMean Left PGEigengap FiberN Right LogSpanningForestN FAMean Left MinCutBalDivSum FiberNDivLength All LogSpanningForestN FiberNDivLength All LogSpanningForestN Unweighted Right Sum Unweighted Left PGEigengap FAMean All LogSpanningForestN FAMean Left Sum FAMean Right LogSpanningForestN FAMean Left PGEigengap FAMean Left PGEigengap Unweighted Left MinCutBalDivSum FiberLengthMean Left Sum FAMean Left MinCutBalDivSum Unweighted Left PGEigengap FiberLengthMean Right LogSpanningForestN FAMean Left PGEigengap Unweighted Left PGEigengap FAMean Left LogSpanningForestN FiberN Left Sum Unweighted Left LogSpanningForestN FAMean Right MinCutBalDivSum Unweighted Right LogSpanningForestN FiberNDivLength Left LogSpanningForestN FAMean Right PGEigengap Unweighted Left Sum FAMean Left AdjLMaxDivD FiberN Right Sum Unweighted Right PGEigengap FiberN All Sum FiberN 11 0.00014 0.00000 0.00029 0.00062 0.00041 0.00378 0.00085 0.00293 0.00015 0.00002 0.00343 0.00113 0.00411 0.00012 0.00019 0.00265 0.00206 0.00382 0.00043 0.00066 0.00143 0.00031 0.00000 0.00218 0.00068 0.00995 0.00019 0.00026 0.00067 0.00141 0.00458 0.00892 0.00056 0.00154 0.00554 0.00380 0.00176 0.00215 0.00012 0.00232 0.00082 0.00462 0.00022 0.000129 0.00095 0.00032 0.00501 0.00224 0.00009 0.00000 0.00047 0.00048 0.00050 0.00051 0.00053 0.00068 0.00084 0.00092 0.00097 0.00108 0.00116 0.00121 0.00123 0.00126 0.00128 0.00134 0.00136 0.00142 0.00150 0.00163 0.00170 0.00175 0.00177 0.00182 0.00186 0.00191 0.00211 0.00212 0.00239 0.00240 0.00243 0.00245 0.00279 0.00289 0.00295 0.00305 0.00338 0.00359 0.00395 0.00456 0.00496 0.00543 0.00587 0.00595 0.00626 0.00660 0.00804 0.00845 0.00910 0.00938 0.04988 0.05045 0.05260 0.05355 0.05414 0.06936 0.08454 0.09212 0.09650 0.10539 0.11274 0.11629 0.11646 0.11823 0.11891 0.12351 0.12370 0.12775 0.13343 0.14369 0.14769 0.15023 0.15009 0.15279 0.15417 0.15694 0.17093 0.16978 0.18842 0.18684 0.18738 0.18596 0.20893 0.21355 0.21516 0.21935 0.24029 0.25152 0.27269 0.31006 0.33212 0.35825 0.38180 0.38054 0.39459 0.40907 0.49040 0.50692 0.53671 0.54418 234 129 83 129 129 129 129 83 234 83 129 129 129 234 83 83 234 83 234 83 129 83 234 234 129 129 129 234 129 83 83 83 234 234 234 83 83 129 129 129 83 234 129 234 83 129 234 83 83 234 Right PGEigengap FAMean Right PGEigengap FAMean Right PGEigengap Unweighted Right MinCutBalDivSum FiberN Right MinCutBalDivSum Unweighted Left LogSpanningForestN FAMean Left LogSpanningForestN FiberN All Sum FiberN All Sum FiberN Right LogSpanningForestN Unweighted Left LogSpanningForestN FiberNDivLength Right PGEigengap Unweighted Right PGEigengap FiberN All MinVertexCover FAMean All HoffmanBound FAMean All Sum FiberNDivLength Right MinCutBalDivSum FiberN Right LogSpanningForestN FiberN Right MinCutBalDivSum FiberLengthMean Right MinCutBalDivSum FiberNDivLength Left MinCutBalDivSum FiberNDivLength Right PGEigengap FAMean All LogSpanningForestN FiberN Right PGEigengap FiberLengthMean Right MinCutBalDivSum FiberLengthMean All Sum FiberNDivLength Right LogSpanningForestN FiberNDivLength All LogSpanningForestN FiberNDivLength Right LogSpanningForestN FiberN Right MinCutBalDivSum FiberN Right HoffmanBound FiberNDivLength Right PGEigengap FiberLengthMean Left MinCutBalDivSum FiberNDivLength Left MinVertexCover FAMean All Sum FiberNDivLength Right Sum FiberN Right Sum FiberNDivLength Left Sum FiberN Right Sum FiberN Left HoffmanBound Unweighted Left Sum FiberN Left Sum FiberN Right Sum FiberNDivLength Right Sum FiberN Left Sum FiberNDivLength Left Sum FiberNDivLength Right Sum FiberNDivLength Right HoffmanBound FAMean All MinVertexCoverBinary Unweighted Right LogSpanningForestN FiberNDivLength 12 0.00074 0.00296 0.00087 0.00563 0.00492 0.00106 0.00014 0.00000 0.00000 0.00541 0.00288 0.00242 0.00869 0.00289 0.00087 0.00002 0.00234 0.00083 0.00234 0.00072 0.00019 0.00112 0.00091 0.00367 0.00768 0.00008 0.00051 0.00106 0.00045 0.00346 0.005129 0.00949 0.00642 0.00107 0.00044 0.00000 0.00018 0.00000 0.00001 0.00848 0.00000 0.00040 0.00180 0.00012 0.00043 0.00100 0.00562 0.00587 0.00716 0.00940 0.00974 0.00981 0.01053 0.01101 0.01212 0.01218 0.01258 0.01290 0.01358 0.01438 0.01447 0.01676 0.01706 0.01713 0.02011 0.02117 0.02197 0.02539 0.02663 0.02854 0.02897 0.02948 0.03308 0.03369 0.04500 0.04728 0.04891 0.05095 0.05578 0.06284 0.06309 0.06515 0.06548 0.07139 0.07318 0.07799 0.07920 0.08380 0.08653 0.08944 0.09430 0.11447 0.12102 0.16411 0.16774 0.22542 0.23691 0.32069 0.38829 0.40996 0.55538 0.54933 0.57889 0.59432 0.64227 0.63359 0.64134 0.64480 0.66520 0.69010 0.67995 0.77084 0.76750 0.75373 0.86462 0.88929 0.90065 1.01567 1.03841 1.08446 1.07195 1.06119 1.15795 1.14542 1.48511 1.51293 1.51627 1.52842 1.61751 1.75951 1.70341 1.69395 1.63696 1.71336 1.68305 1.71586 1.66329 1.67598 1.64406 1.60984 1.60310 1.83157 1.81523 2.29752 2.18062 2.70502 2.60604 3.20692 3.49459 3.27971 83 83 83 234 129 83 234 Left HoffmanBound FiberN All MinVertexCover FiberNDivLength Right MinSpanningForest FiberLengthMean Right MinSpanningForest FiberLengthMean All MinVertexCover FiberN All MinVertexCover FiberN All MinVertexCover FiberN 0.00175 0.00036 0.00491 0.00601 0.00232 0.00244 0.00055 0.41913 0.46677 0.55239 0.55631 0.71406 0.84437 0.92958 2.93394 2.80065 2.76195 2.22523 2.14217 1.68874 0.92958 Table 1: The results and the statistical analysis of the graph-theoretical evaluation of the sex differences in the 96 diffusion MRI images. The first column gives the resolution in each hemisphere; the number of nodes in the whole graph is 83, 129 and 234, respectively. The second column describes the graph parameter computed: its syntactics is as follows: each parameter-name contains two separating " " symbols that define three parts of the parameter-name. The first part describe the hemi- sphere or the whole connectome with the words Left, Right or All. The second part describes the parameter computed, and the third part the weight function used (their definitions are given in section "Materials and methods"). The third column contains the p-values of the first round, the second column the p-values of the second round, and the third col- umn the (very strict) Holm-Bonferroni correction of the p-value. With p=0.05 all the first 12 rows describe significantly different graph theo- retical properties between sexes. One-by-one, each row with italic third column describe significant differences between sexes, with p=0.05. For the details we refer to the section "Statistical analysis". 6. Acknowledgments The authors declare no conflicts of interest. References [1] P. Hagmann, P. E. Grant, D. A. Fair, Mr connectomics: a conceptual framework for studying the developing brain., Front Syst Neurosci 6 (2012) 43. doi:10.3389/fnsys.2012.00043. URL http://dx.doi.org/10.3389/fnsys.2012.00043 [2] R. C. Craddock, M. P. Milham, S. M. LaConte, Predicting intrinsic brain activity., Neuroimage 82 (2013) 127–136. doi:10.1016/j.neuroimage. 2013.05.072. URL http://dx.doi.org/10.1016/j.neuroimage.2013.05.072 [3] G. Ball, P. Aljabar, S. Zebari, N. Tusor, T. Arichi, N. Merchant, E. C. Robinson, E. Ogundipe, D. Rueckert, A. D. Edwards, S. J. Counsell, Rich- club organization of the newborn human brain., Proc Natl Acad Sci U S A 111 (20) (2014) 7456–7461. doi:10.1073/pnas.1324118111. URL http://dx.doi.org/10.1073/pnas.1324118111 13 [4] C. I. Bargmann, Beyond the connectome: how neuromodulators shape neural circuits., Bioessays 34 (6) (2012) 458–465. doi:10.1002/bies. 201100185. URL http://dx.doi.org/10.1002/bies.201100185 [5] D. Batalle, E. Munoz-Moreno, F. Figueras, N. Bargallo, E. Eixarch, E. Gratacos, Normalization of similarity-based individual brain networks from gray matter MRI and its association with neurodevelopment in in- fants with intrauterine growth restriction., Neuroimage 83 (2013) 901–911. doi:10.1016/j.neuroimage.2013.07.045. URL http://dx.doi.org/10.1016/j.neuroimage.2013.07.045 [6] D. J. Graham, Routing in the brain., Front Comput Neurosci 8 (2014) 44. doi:10.3389/fncom.2014.00044. URL http://dx.doi.org/10.3389/fncom.2014.00044 [7] F. Agosta, S. Galantucci, P. Valsasina, E. Canu, A. Meani, A. Marcone, G. Magnani, A. Falini, G. Comi, M. Filippi, Disrupted brain connectome in semantic variant of primary progressive aphasia., Neurobiol Agingdoi: 10.1016/j.neurobiolaging.2014.05.017. URL http://dx.doi.org/10.1016/j.neurobiolaging.2014.05.017 [8] A. F. Alexander-Bloch, P. T. Reiss, J. Rapoport, H. McAdams, J. N. Giedd, E. T. Bullmore, N. Gogtay, Abnormal cortical growth in schizophrenia targets normative modules of synchronized development., Biol Psychia- trydoi:10.1016/j.biopsych.2014.02.010. URL http://dx.doi.org/10.1016/j.biopsych.2014.02.010 [9] J. T. Baker, A. J. Holmes, G. A. Masters, B. T. T. Yeo, F. Krienen, R. L. Buckner, D. Ongur, Disruption of cortical association networks in schizophrenia and psychotic bipolar disorder., JAMA Psychiatry 71 (2) (2014) 109–118. doi:10.1001/jamapsychiatry.2013.3469. URL http://dx.doi.org/10.1001/jamapsychiatry.2013.3469 [10] P. Besson, V. Dinkelacker, R. Valabregue, L. Thivard, X. Leclerc, M. Baulac, D. Sammler, O. Colliot, S. Leh´ericy, S. Samson, S. Dupont, Structural connectivity differences in left and right temporal lobe epilepsy., Neuroimage 100C (2014) 135–144. doi:10.1016/j.neuroimage.2014.04. 071. URL http://dx.doi.org/10.1016/j.neuroimage.2014.04.071 [11] L. Bonilha, T. Nesland, C. Rorden, P. Fillmore, R. P. Ratnayake, J. Fridriksson, Mapping remote subcortical ramifications of injury after ischemic strokes., Behav Neurol 2014 (2014) 215380. doi:10.1155/2014/ 215380. URL http://dx.doi.org/10.1155/2014/215380 14 [12] M. Ingalhalikar, A. Smith, D. Parker, T. D. Satterthwaite, M. A. Elliott, K. Ruparel, H. Hakonarson, R. E. Gur, R. C. Gur, R. Verma, Sex differ- ences in the structural connectome of the human brain., Proc Natl Acad Sci U S A 111 (2) (2014) 823–828. doi:10.1073/pnas.1316909110. URL http://dx.doi.org/10.1073/pnas.1316909110 [13] D. Joel, R. Tarrasch, On the mis-presentation and misinterpretation of gender-related data: the case of Ingalhalikar's human connectome study., Proc Natl Acad Sci U S A 111 (6) (2014) E637. doi:10.1073/pnas. 1323319111. URL http://dx.doi.org/10.1073/pnas.1323319111 [14] M. Ingalhalikar, A. Smith, D. Parker, T. D. Satterthwaite, M. A. Elliott, K. Ruparel, H. Hakonarson, R. E. Gur, R. C. Gur, R. Verma, Reply to Joel and Tarrasch: On misreading and shooting the messenger., Proc Natl Acad Sci U S A 111 (6) (2014) E638. [15] C. Fine, Neuroscience. his brain, her brain?, Science 346 (6212) (2014) 915– 916. doi:10.1126/science.1262061. URL http://dx.doi.org/10.1126/science.1262061 [16] S. F. Witelson, H. Beresh, D. L. Kigar, Intelligence and brain size in 100 postmortem brains: sex, lateralization and age factors., Brain 129 (Pt 2) (2006) 386–398. doi:10.1093/brain/awh696. URL http://dx.doi.org/10.1093/brain/awh696 [17] Y. Taki, B. Thyreau, S. Kinomura, K. Sato, R. Goto, R. Kawashima, H. Fukuda, Correlations among brain gray matter volumes, age, gender, and hemisphere in healthy individuals., PLoS One 6 (7) (2011) e22734. doi:10.1371/journal.pone.0022734. URL http://dx.doi.org/10.1371/journal.pone.0022734 [18] E. L. Lawler, Combinatorial optimization: networks and matroids, Courier Dover Publications, 1976. [19] L. R. Ford, D. R. Fulkerson, Maximal flow through a network, Canadian Journal of Mathematics 8 (3) (1956) 399–404. [20] P. Ade, N. Aghanim, C. Armitage-Caplan, M. Arnaud, M. Ashdown, F. Atrio-Barandela, J. Aumont, C. Baccigalupi, A. Banday, R. Barreiro, et al., Planck 2013 results. XVI. Cosmological parameters, arXiv preprint arXiv:1303.5076. [21] M. R. Garey, D. S. Johnson, L. Stockmeyer, Some simplified NP-complete graph problems, Theoretical computer science 1 (3) (1976) 237–267. [22] R. E. Tarjan, Data structures and network algorithms, Vol. 44 of CBMS- NSF Regional Conference Series in Applied Mathematics, Society for In- dustrial Applied Mathematics, 1983. 15 [23] S. Hoory, N. Linial, A. Wigderson, Expander graphs and their applications, Bulletin of the American Mathematical Society 43 (4) (2006) 439–561. [24] L. Lov´asz, Combinatorial problems and exercises, 2nd Edition, American Mathematical Society, 2007. [25] H. Prufer, Neuer Beweis eines Satzes uber Permutationen, Arch. Math. Phys 27 (1918) 742–744. [26] G. Kirchhoff, uber die Auflosung der Gleichungen, auf welche man bei der untersuchung der linearen verteilung galvanischer Strome gefuhrt wird, Ann. Phys. Chem. 72. [27] F. R. Chung, Spectral graph theory, Vol. 92, American Mathematical Soc., 1997. [28] J. A. McNab, B. L. Edlow, T. Witzel, S. Y. Huang, H. Bhat, K. Heberlein, T. Feiweier, K. Liu, B. Keil, J. Cohen-Adad, M. D. Tisdall, R. D. Folkerth, H. C. Kinney, L. L. Wald, The Human Connectome Project and beyond: initial applications of 300 mT/m gradients., Neuroimage 80 (2013) 234– 245. doi:10.1016/j.neuroimage.2013.05.074. URL http://dx.doi.org/10.1016/j.neuroimage.2013.05.074 [29] A. Daducci, S. Gerhard, A. Griffa, A. Lemkaddem, L. Cammoun, X. Gi- gandet, R. Meuli, P. Hagmann, J.-P. Thiran, The connectome mapper: an open-source processing pipeline to map connectomes with MRI., PLoS One 7 (12) (2012) e48121. doi:10.1371/journal.pone.0048121. URL http://dx.doi.org/10.1371/journal.pone.0048121 [30] P. J. Basser, C. Pierpaoli, Microstructural and physiological features of tissues elucidated by quantitative-diffusion-tensor mri. 1996., J Magn Reson 213 (2) (2011) 560–570. doi:10.1016/j.jmr.2011.09.022. URL http://dx.doi.org/10.1016/j.jmr.2011.09.022 [31] D. S. Hochbaum, Approximation algorithms for the set covering and vertex cover problems, SIAM Journal on Computing 11 (3) (1982) 555–556. [32] T. Achterberg, T. Berthold, T. Koch, K. Wolter, Constraint integer pro- gramming: A new approach to integrate CP and MIP, in: Integration of AI and OR techniques in constraint programming for combinatorial opti- mization problems, Springer, 2008, pp. 6–20. [33] T. Achterberg, Scip: solving constraint integer programs, Mathematical Programming Computation 1 (1) (2009) 1–41. [34] P. G. Hoel, Introduction to mathematical statistics., 5th Edition, John Wiley & Sons, Inc., New York, 1984. [35] T. H. Wonnacott, R. J. Wonnacott, Introductory statistics, Vol. 19690, Wiley New York, 1972. 16 [36] S. Holm, A simple sequentially rejective multiple test procedure, Scandi- navian Journal of Statistics (1979) 65–70. 17 Appendix In this appendix we list the graph-theoretic parameters computed for the resolutions of 83, 129 and 234 vertex graphs. The tables contain their arith- metic means in the male and female groups, and the corresponding p-values. The values in these tables contain the values corresponded to round 1 (see the "Statistical analysis" subsection in the main text). The graph-parameters are defined in the caption of Table 1. Significant differences (p < 0.01) are denoted with an asterisk in the last column. Scale 83, round 1 Property All AdjLMaxDivD FAMean All AdjLMaxDivD FiberLengthMean All AdjLMaxDivD FiberN All AdjLMaxDivD FiberNDivLength All AdjLMaxDivD Unweighted All HoffmanBound FAMean All HoffmanBound FiberLengthMean All HoffmanBound FiberN All HoffmanBound FiberNDivLength All HoffmanBound Unweighted All LogSpanningForestN FAMean All LogSpanningForestN FiberLengthMean All LogSpanningForestN FiberN All LogSpanningForestN FiberNDivLength All LogSpanningForestN Unweighted All MinCutBalDivSum FAMean All MinCutBalDivSum FiberLengthMean All MinCutBalDivSum FiberN All MinCutBalDivSum FiberNDivLength All MinCutBalDivSum Unweighted All MinSpanningForest FAMean All MinSpanningForest FiberLengthMean All MinSpanningForest FiberN All MinSpanningForest FiberNDivLength All MinVertexCoverBinary Unweighted All MinVertexCover FAMean All MinVertexCover FiberLengthMean All MinVertexCover FiberN All MinVertexCover FiberNDivLength All MinVertexCover Unweighted All PGEigengap FAMean All PGEigengap FiberLengthMean All PGEigengap FiberN All PGEigengap FiberNDivLength All PGEigengap Unweighted All Sum FAMean All Sum FiberLengthMean All Sum FiberN All Sum FiberNDivLength All Sum Unweighted Left AdjLMaxDivD FAMean Left AdjLMaxDivD FiberLengthMean Left AdjLMaxDivD FiberN Left AdjLMaxDivD FiberNDivLength Left AdjLMaxDivD Unweighted Left HoffmanBound FAMean Left HoffmanBound FiberLengthMean Left HoffmanBound FiberN Female 1.36008 1.44214 2.02416 1.84476 1.26760 4.36096 3.21938 2.63525 2.51038 4.55192 110.69890 456.60084 397.53780 148.03174 191.66035 0.00793 0.03115 0.02924 0.02868 0.04001 19.78188 1096.37958 99.53846 3.65548 59.80769 18.73144 2014.06431 2427.21154 110.25657 40.90385 0.05403 0.04167 0.03156 0.03470 0.05214 222.01291 Male 1.37750 1.43602 2.10529 1.86864 1.26456 4.18564 3.26552 2.55573 2.40550 4.43931 101.82758 452.95875 389.79037 139.85355 187.85180 0.00474 0.02889 0.02711 0.02644 0.03721 18.63722 1112.97289 102.93333 3.66822 59.00000 18.10619 1955.70824 2315.20000 103.59777 41.00000 0.05071 0.03891 0.02829 0.03062 0.04740 201.02562 16845.33062 11261.65385 476.56342 567.07692 15792.24352 10237.13333 433.37987 539.80000 1.33644 1.40515 1.90607 1.71498 1.24027 4.55406 3.25098 2.71430 1.35216 1.38890 2.02087 1.77482 1.23523 4.38621 3.28435 2.61098 p-value 0.06806 0.72030 0.05606 0.41834 0.63251 0.00087 0.33136 0.03144 0.01815 0.04616 0.00012 0.18687 0.00001 0.00003 0.00113 0.14869 0.47008 0.34092 0.38768 0.28887 0.02232 0.10506 0.14280 0.93669 0.00716 0.01699 0.37460 0.00244 0.00036 0.32897 0.28914 0.43309 0.03885 0.01847 0.09708 0.00029 0.06219 0.00000 0.00002 0.00025 0.15767 0.32795 0.00501 0.07539 0.43598 0.01297 0.51250 0.00175 ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ 18 Left HoffmanBound FiberNDivLength Left HoffmanBound Unweighted Left LogSpanningForestN FAMean Left LogSpanningForestN FiberLengthMean Left LogSpanningForestN FiberN Left LogSpanningForestN FiberNDivLength Left LogSpanningForestN Unweighted Left MinCutBalDivSum FAMean Left MinCutBalDivSum FiberLengthMean Left MinCutBalDivSum FiberN Left MinCutBalDivSum FiberNDivLength Left MinCutBalDivSum Unweighted Left MinSpanningForest FAMean Left MinSpanningForest FiberLengthMean Left MinSpanningForest FiberN Left MinSpanningForest FiberNDivLength Left MinVertexCoverBinary Unweighted Left MinVertexCover FAMean Left MinVertexCover FiberLengthMean Left MinVertexCover FiberN Left MinVertexCover FiberNDivLength Left MinVertexCover Unweighted Left PGEigengap FAMean Left PGEigengap FiberLengthMean Left PGEigengap FiberN Left PGEigengap FiberNDivLength Left PGEigengap Unweighted Left Sum FAMean Left Sum FiberLengthMean Left Sum FiberN Left Sum FiberNDivLength Left Sum Unweighted Right AdjLMaxDivD FAMean Right AdjLMaxDivD FiberLengthMean Right AdjLMaxDivD FiberN Right AdjLMaxDivD FiberNDivLength Right AdjLMaxDivD Unweighted Right HoffmanBound FAMean Right HoffmanBound FiberLengthMean Right HoffmanBound FiberN Right HoffmanBound FiberNDivLength Right HoffmanBound Unweighted Right LogSpanningForestN FAMean Right LogSpanningForestN FiberLengthMean Right LogSpanningForestN FiberN Right LogSpanningForestN FiberNDivLength Right LogSpanningForestN Unweighted Right MinCutBalDivSum FAMean Right MinCutBalDivSum FiberLengthMean Right MinCutBalDivSum FiberN Right MinCutBalDivSum FiberNDivLength Right MinCutBalDivSum Unweighted Right MinSpanningForest FAMean Right MinSpanningForest FiberLengthMean Right MinSpanningForest FiberN Right MinSpanningForest FiberNDivLength Right MinVertexCoverBinary Unweighted Right MinVertexCover FAMean Right MinVertexCover FiberLengthMean Right MinVertexCover FiberN Right MinVertexCover FiberNDivLength Right MinVertexCover Unweighted Right PGEigengap FAMean Right PGEigengap FiberLengthMean Right PGEigengap FiberN Right PGEigengap FiberNDivLength Right PGEigengap Unweighted Right Sum FAMean 2.66652 4.73205 53.30579 229.63370 199.27958 73.53683 95.46307 0.00687 0.23438 0.13337 0.11057 0.24513 9.57924 561.47024 51.23077 1.82447 30.23077 9.23616 2.59451 4.57434 48.82905 227.32675 195.25428 69.82889 93.39767 0.00320 0.21147 0.12011 0.09321 0.22019 9.06313 560.36391 53.73333 1.89521 29.73333 8.88642 1064.27185 1158.21154 1027.73430 1143.46667 54.26322 20.80769 0.33446 0.33383 0.16980 0.14486 0.30646 106.64056 8629.73791 5514.61538 233.06402 282.50000 1.32878 1.39672 2.00803 1.76990 1.25268 4.47438 3.33823 2.67311 2.62635 4.61480 52.25642 218.25106 190.62427 69.84080 90.24090 0.02476 0.24577 0.13346 0.10831 0.23713 10.30911 532.13580 50.76923 1.94340 29.07692 9.26572 51.17634 20.83333 0.29469 0.29287 0.15238 0.13413 0.27160 96.80731 8122.82646 5049.73333 213.49323 269.06667 1.34242 1.38478 2.09048 1.81343 1.24720 4.28666 3.39478 2.57701 2.48983 4.50726 48.14346 216.24411 187.02757 66.17446 88.51678 0.00851 0.22309 0.12050 0.09357 0.22022 9.79708 547.85331 52.53333 1.89232 28.73333 9.03965 934.26071 1169.63462 897.95882 1122.93333 53.57144 20.11538 0.32454 0.34029 0.17666 0.15245 0.29582 105.62164 51.50298 20.26667 0.28808 0.29461 0.15912 0.14034 0.26081 95.26436 0.13782 0.01379 0.00082 0.18765 0.00012 0.00343 0.01389 0.17151 0.01779 0.00403 0.00031 0.00206 0.04242 0.87722 0.26795 0.62729 0.09601 0.01371 0.35926 0.55321 0.02122 0.75017 0.00215 0.01329 0.01654 0.02837 0.00458 0.00056 0.13250 0.00000 0.00043 0.00378 0.14511 0.30191 0.05380 0.09784 0.29540 0.00587 0.29902 0.05411 0.00560 0.03806 0.00067 0.16431 0.00083 0.00022 0.00541 0.00496 0.02216 0.00346 0.00072 0.01629 0.10419 0.00491 0.26282 0.58863 0.15457 0.12382 0.23661 0.07986 0.10452 0.10527 0.00112 0.00949 0.02617 0.01613 0.00087 0.00028 ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ 19 Right Sum FiberLengthMean Right Sum FiberN Right Sum FiberNDivLength Right Sum Unweighted 7644.90330 5378.03846 225.94776 261.30769 7086.91000 4884.66667 206.97587 248.26667 0.02974 0.00000 0.00018 0.00019 ∗ ∗ ∗ Scale 129, round 1 Property All AdjLMaxDivD FAMean All AdjLMaxDivD FiberLengthMean All AdjLMaxDivD FiberN All AdjLMaxDivD FiberNDivLength All AdjLMaxDivD Unweighted All HoffmanBound FAMean All HoffmanBound FiberLengthMean All HoffmanBound FiberN All HoffmanBound FiberNDivLength All HoffmanBound Unweighted All LogSpanningForestN FAMean All LogSpanningForestN FiberLengthMean All LogSpanningForestN FiberN All LogSpanningForestN FiberNDivLength All LogSpanningForestN Unweighted All MinCutBalDivSum FAMean All MinCutBalDivSum FiberLengthMean All MinCutBalDivSum FiberN All MinCutBalDivSum FiberNDivLength All MinCutBalDivSum Unweighted All MinSpanningForest FAMean All MinSpanningForest FiberLengthMean All MinSpanningForest FiberN All MinSpanningForest FiberNDivLength All MinVertexCoverBinary Unweighted All MinVertexCover FAMean All MinVertexCover FiberLengthMean All MinVertexCover FiberN All MinVertexCover FiberNDivLength All MinVertexCover Unweighted All PGEigengap FAMean All PGEigengap FiberLengthMean All PGEigengap FiberN All PGEigengap FiberNDivLength All PGEigengap Unweighted All Sum FAMean All Sum FiberLengthMean All Sum FiberN All Sum FiberNDivLength All Sum Unweighted Left AdjLMaxDivD FAMean Left AdjLMaxDivD FiberLengthMean Left AdjLMaxDivD FiberN Left AdjLMaxDivD FiberNDivLength Left AdjLMaxDivD Unweighted Left HoffmanBound FAMean Left HoffmanBound FiberLengthMean Left HoffmanBound FiberN Left HoffmanBound FiberNDivLength Left HoffmanBound Unweighted Left LogSpanningForestN FAMean Left LogSpanningForestN FiberLengthMean Left LogSpanningForestN FiberN Left LogSpanningForestN FiberNDivLength Left LogSpanningForestN Unweighted Left MinCutBalDivSum FAMean Left MinCutBalDivSum FiberLengthMean Left MinCutBalDivSum FiberN Left MinCutBalDivSum FiberNDivLength Female 1.40519 1.50483 2.14552 2.09783 1.30028 4.40157 3.19684 2.50604 2.34647 4.62935 194.37749 739.78985 599.76631 210.52236 322.09324 0.00668 0.01706 0.02658 0.02495 0.02218 30.14746 1642.68263 140.23077 4.42401 96.46154 29.56250 3230.07900 2444.92308 120.18766 63.88462 0.03143 0.02427 0.02781 0.02880 0.03012 397.68878 30670.09535 12375.61538 548.61301 1020.80769 1.37823 1.43638 1.84672 1.77313 1.26380 4.57539 3.23550 2.80373 2.70077 4.75280 96.11000 373.09476 300.77613 105.01323 162.01302 0.00873 0.19822 0.12848 0.06926 Male 1.42604 1.50158 2.22254 2.04782 1.29097 4.29660 3.24689 2.48884 2.41938 4.51267 181.03525 732.55388 588.61699 200.75240 316.62672 0.00324 0.01607 0.02429 0.02258 0.02065 28.58509 1664.23693 140.93333 4.43795 96.26667 28.72424 3121.21684 2337.40000 116.22553 63.96667 0.02928 0.02260 0.02453 0.02498 0.02725 360.50850 28478.19852 11458.13333 510.71378 972.86667 1.39812 1.42179 1.92762 1.80979 1.25501 4.44885 3.25088 2.74220 2.64308 4.61941 89.25516 368.65582 295.83044 100.80980 158.88026 0.00273 0.17378 0.10467 0.05546 p-value 0.10040 0.87806 0.15242 0.32031 0.27278 0.02644 0.32568 0.64956 0.07720 0.01233 0.00019 0.09867 0.00000 0.00000 0.00218 0.05930 0.56293 0.26627 0.30029 0.30082 0.02073 0.07510 0.55077 0.92181 0.66793 0.02181 0.29100 0.00232 0.02502 0.35805 0.25524 0.43054 0.01902 0.00792 0.09661 0.00015 0.03582 0.00000 0.00008 0.00026 0.12792 0.36739 0.12247 0.33521 0.16858 0.01512 0.77158 0.14090 0.21782 0.00848 0.00106 0.08843 0.00014 0.00288 0.01336 0.05683 0.00892 0.00001 0.00019 ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ 20 Left MinCutBalDivSum Unweighted Left MinSpanningForest FAMean Left MinSpanningForest FiberLengthMean Left MinSpanningForest FiberN Left MinSpanningForest FiberNDivLength Left MinVertexCoverBinary Unweighted Left MinVertexCover FAMean Left MinVertexCover FiberLengthMean Left MinVertexCover FiberN Left MinVertexCover FiberNDivLength Left MinVertexCover Unweighted Left PGEigengap FAMean Left PGEigengap FiberLengthMean Left PGEigengap FiberN Left PGEigengap FiberNDivLength Left PGEigengap Unweighted Left Sum FAMean Left Sum FiberLengthMean Left Sum FiberN Left Sum FiberNDivLength Left Sum Unweighted Right AdjLMaxDivD FAMean Right AdjLMaxDivD FiberLengthMean Right AdjLMaxDivD FiberN Right AdjLMaxDivD FiberNDivLength Right AdjLMaxDivD Unweighted Right HoffmanBound FAMean Right HoffmanBound FiberLengthMean Right HoffmanBound FiberN Right HoffmanBound FiberNDivLength Right HoffmanBound Unweighted Right LogSpanningForestN FAMean Right LogSpanningForestN FiberLengthMean Right LogSpanningForestN FiberN Right LogSpanningForestN FiberNDivLength Right LogSpanningForestN Unweighted Right MinCutBalDivSum FAMean Right MinCutBalDivSum FiberLengthMean Right MinCutBalDivSum FiberN Right MinCutBalDivSum FiberNDivLength Right MinCutBalDivSum Unweighted Right MinSpanningForest FAMean Right MinSpanningForest FiberLengthMean Right MinSpanningForest FiberN Right MinSpanningForest FiberNDivLength Right MinVertexCoverBinary Unweighted Right MinVertexCover FAMean Right MinVertexCover FiberLengthMean Right MinVertexCover FiberN Right MinVertexCover FiberNDivLength Right MinVertexCover Unweighted Right PGEigengap FAMean Right PGEigengap FiberLengthMean Right PGEigengap FiberN Right PGEigengap FiberNDivLength Right PGEigengap Unweighted Right Sum FAMean Right Sum FiberLengthMean Right Sum FiberN Right Sum FiberNDivLength Right Sum Unweighted 0.19535 14.57467 828.34729 69.30769 2.16989 48.76923 14.65360 0.17339 13.88500 834.54850 72.20000 2.25626 48.86667 14.09857 1700.29684 1169.82692 1637.18742 1125.20000 58.76113 32.28846 0.22611 0.23241 0.12346 0.09689 0.20204 56.23736 32.30000 0.19656 0.20065 0.10569 0.08572 0.17516 197.41850 16079.40944 6071.96154 269.09760 519.53846 178.80563 14931.40760 5641.93333 251.40080 492.86667 1.35746 1.42015 2.05564 1.82146 1.26684 4.37886 3.32686 2.66511 2.68679 4.60861 93.41904 358.00491 291.08563 100.74383 154.36558 0.02361 0.20000 0.11452 0.06865 0.19180 15.61479 808.14079 70.46154 2.32813 47.34615 14.70648 1.36837 1.41129 2.19134 1.86716 1.25522 4.29574 3.36662 2.56838 2.59830 4.51407 87.28295 354.73456 285.72242 96.22891 151.96595 0.01005 0.17303 0.10111 0.06326 0.16911 14.88977 824.37649 68.93333 2.26810 47.00000 14.40974 1516.99670 1175.50000 1461.52391 1166.36667 59.59421 31.61538 0.22838 0.23840 0.12500 0.10075 0.20584 58.78162 31.73333 0.19627 0.19868 0.11049 0.09371 0.17429 190.48228 13952.01182 5935.73077 262.31420 477.38462 172.48988 13003.32443 5525.26667 246.32048 454.86667 0.00265 0.06189 0.36946 0.02902 0.53695 0.69355 0.01273 0.30481 0.06266 0.06303 0.88865 0.00995 0.02197 0.00382 0.00223 0.01081 0.00032 0.07487 0.00000 0.00100 0.00232 0.36353 0.54264 0.01338 0.20816 0.12057 0.20294 0.49418 0.01727 0.01992 0.08448 0.00143 0.14280 0.00045 0.00051 0.01158 0.00807 0.00768 0.00563 0.09375 0.00492 0.06537 0.03729 0.07096 0.46298 0.29760 0.13709 0.23679 0.68666 0.47843 0.20363 0.00296 0.01013 0.00869 0.03033 0.00242 0.00062 0.04620 0.00001 0.00180 0.00068 ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ Scale 234, round 1 Property All AdjLMaxDivD FAMean All AdjLMaxDivD FiberLengthMean Female 2.15050 2.35868 Male 2.14489 2.34695 p-value 0.86385 0.80876 21 All AdjLMaxDivD FiberN All AdjLMaxDivD FiberNDivLength All AdjLMaxDivD Unweighted All HoffmanBound FAMean All HoffmanBound FiberLengthMean All HoffmanBound FiberN All HoffmanBound FiberNDivLength All HoffmanBound Unweighted All LogSpanningForestN FAMean All LogSpanningForestN FiberLengthMean All LogSpanningForestN FiberN All LogSpanningForestN FiberNDivLength All LogSpanningForestN Unweighted All MinCutBalDivSum FAMean All MinCutBalDivSum FiberLengthMean All MinCutBalDivSum FiberN All MinCutBalDivSum FiberNDivLength All MinCutBalDivSum Unweighted All MinSpanningForest FAMean All MinSpanningForest FiberLengthMean All MinSpanningForest FiberN All MinSpanningForest FiberNDivLength All MinVertexCoverBinary Unweighted All MinVertexCover FAMean All MinVertexCover FiberLengthMean All MinVertexCover FiberN All MinVertexCover FiberNDivLength All MinVertexCover Unweighted All PGEigengap FAMean All PGEigengap FiberLengthMean All PGEigengap FiberN All PGEigengap FiberNDivLength All PGEigengap Unweighted All Sum FAMean All Sum FiberLengthMean All Sum FiberN All Sum FiberNDivLength All Sum Unweighted Left AdjLMaxDivD FAMean Left AdjLMaxDivD FiberLengthMean Left AdjLMaxDivD FiberN Left AdjLMaxDivD FiberNDivLength Left AdjLMaxDivD Unweighted Left HoffmanBound FAMean Left HoffmanBound FiberLengthMean Left HoffmanBound FiberN Left HoffmanBound FiberNDivLength Left HoffmanBound Unweighted Left LogSpanningForestN FAMean Left LogSpanningForestN FiberLengthMean Left LogSpanningForestN FiberN Left LogSpanningForestN FiberNDivLength Left LogSpanningForestN Unweighted Left MinCutBalDivSum FAMean Left MinCutBalDivSum FiberLengthMean Left MinCutBalDivSum FiberN Left MinCutBalDivSum FiberNDivLength Left MinCutBalDivSum Unweighted Left MinSpanningForest FAMean Left MinSpanningForest FiberLengthMean Left MinSpanningForest FiberN Left MinSpanningForest FiberNDivLength Left MinVertexCoverBinary Unweighted Left MinVertexCover FAMean Left MinVertexCover FiberLengthMean Left MinVertexCover FiberN Left MinVertexCover FiberNDivLength Left MinVertexCover Unweighted 5.14838 5.17072 1.89062 3.63940 2.92490 2.23619 2.20178 3.73661 446.86116 2324.68381 1456.24015 149.01647 942.01654 0.00000 0.00769 0.02405 0.00000 0.00898 98.19730 5358.83904 481.46154 18.53246 276.15385 89.53747 8136.04292 2430.61538 129.82332 222.57692 0.01106 0.00860 0.01894 0.01773 0.00995 5.00652 4.78287 1.84578 3.62013 2.98466 2.26557 2.23871 3.72935 416.54482 2325.52712 1445.53700 138.16817 944.27877 0.00000 0.00723 0.02168 0.00000 0.00834 92.47667 5379.38212 479.20000 18.36575 280.33333 87.25805 7957.20990 2344.50000 126.64639 223.33333 0.01201 0.00960 0.01927 0.01767 0.01067 1033.36931 74747.99556 13609.34615 652.17760 2801.69231 961.08503 71461.78993 12823.40000 623.38731 2746.20000 2.14627 2.29338 4.03186 3.93717 1.86339 3.74670 2.94312 2.51168 2.44470 3.82814 212.18613 1159.44274 723.10349 70.44766 467.24325 0.00000 0.09355 0.07158 0.00000 0.09416 47.28302 2702.23206 244.11538 9.45842 137.19231 43.50481 4136.87086 1168.19231 63.94002 111.38462 2.14335 2.29214 4.16381 3.84897 1.81508 3.77335 2.99233 2.47461 2.45140 3.84621 197.08273 1165.33847 723.01322 65.77187 470.94213 0.00000 0.07667 0.05914 0.00000 0.07896 44.78250 2712.65026 244.46667 9.50259 140.00000 42.59720 4052.71473 1153.66667 64.04107 112.26667 0.35870 0.02543 0.06482 0.57408 0.17340 0.30055 0.13550 0.82472 0.03232 0.96824 0.36683 0.15229 0.83734 nan 0.57442 0.21132 0.45008 0.32475 0.00151 0.44199 0.45787 0.71037 0.12225 0.06974 0.48358 0.00056 0.02087 0.39844 0.54543 0.45409 0.89543 0.97772 0.59117 0.00297 0.18467 0.00000 0.00139 0.21290 0.93401 0.97718 0.29128 0.38654 0.04174 0.55549 0.25660 0.28318 0.85286 0.65499 0.04326 0.58696 0.98899 0.31060 0.52729 nan 0.00655 0.00062 nan 0.00153 0.00239 0.49327 0.89014 0.88229 0.06105 0.16942 0.55895 0.46021 0.92511 0.09259 ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ 22 Left PGEigengap FAMean Left PGEigengap FiberLengthMean Left PGEigengap FiberN Left PGEigengap FiberNDivLength Left PGEigengap Unweighted Left Sum FAMean Left Sum FiberLengthMean Left Sum FiberN Left Sum FiberNDivLength Left Sum Unweighted Right AdjLMaxDivD FAMean Right AdjLMaxDivD FiberLengthMean Right AdjLMaxDivD FiberN Right AdjLMaxDivD FiberNDivLength Right AdjLMaxDivD Unweighted Right HoffmanBound FAMean Right HoffmanBound FiberLengthMean Right HoffmanBound FiberN Right HoffmanBound FiberNDivLength Right HoffmanBound Unweighted Right LogSpanningForestN FAMean Right LogSpanningForestN FiberLengthMean Right LogSpanningForestN FiberN Right LogSpanningForestN FiberNDivLength Right LogSpanningForestN Unweighted Right MinCutBalDivSum FAMean Right MinCutBalDivSum FiberLengthMean Right MinCutBalDivSum FiberN Right MinCutBalDivSum FiberNDivLength Right MinCutBalDivSum Unweighted Right MinSpanningForest FAMean Right MinSpanningForest FiberLengthMean Right MinSpanningForest FiberN Right MinSpanningForest FiberNDivLength Right MinVertexCoverBinary Unweighted Right MinVertexCover FAMean Right MinVertexCover FiberLengthMean Right MinVertexCover FiberN Right MinVertexCover FiberNDivLength Right MinVertexCover Unweighted Right PGEigengap FAMean Right PGEigengap FiberLengthMean Right PGEigengap FiberN Right PGEigengap FiberNDivLength Right PGEigengap Unweighted Right Sum FAMean Right Sum FiberLengthMean Right Sum FiberN Right Sum FiberNDivLength Right Sum Unweighted 0.08402 0.08669 0.06812 0.05084 0.07190 0.07554 0.07722 0.05737 0.04481 0.06398 504.02280 38178.70022 6716.53846 322.55630 1401.80769 470.30921 36255.83071 6389.20000 311.23280 1380.33333 2.00996 2.15381 4.11898 3.79534 1.79189 3.63008 3.00591 2.40837 2.45857 3.71704 228.90719 1154.04516 724.05083 72.92465 467.61765 0.00050 0.10021 0.07599 0.00034 0.09573 50.98056 2655.83115 238.96154 9.28191 138.30769 45.80119 3994.00115 1144.80769 62.35579 111.09615 0.08312 0.08538 0.06631 0.05084 0.07102 2.02718 2.18170 4.41926 3.75488 1.77141 3.59884 3.02300 2.33314 2.38848 3.69299 215.28259 1148.91122 716.03208 68.45678 466.56728 0.00000 0.08439 0.06701 0.00000 0.08171 47.79220 2655.71544 236.00000 9.00082 140.00000 44.57707 3884.90036 1129.73333 61.50301 111.10000 0.07683 0.07887 0.06080 0.04854 0.06430 517.36095 35857.03890 6524.53846 312.50248 1368.00000 481.68012 34486.76733 6187.46667 299.09835 1339.06667 0.28777 0.29463 0.09675 0.18106 0.24844 0.01077 0.19037 0.00107 0.04079 0.39428 0.61502 0.41400 0.03397 0.70781 0.38704 0.45778 0.69490 0.00150 0.01602 0.50645 0.07936 0.63377 0.22608 0.30478 0.85195 0.19303 0.01271 0.00641 0.18042 0.01034 0.00435 0.99483 0.15420 0.18645 0.25603 0.07765 0.36802 0.41752 0.47854 0.99385 0.33378 0.40909 0.28067 0.52890 0.25554 0.00745 0.26347 0.00050 0.01170 0.20464 ∗ ∗ ∗ ∗ ∗ ∗ 23
1910.08423
1
1910
2019-10-17T15:03:17
Quantum effects in the brain: A review
[ "q-bio.NC", "quant-ph" ]
In the mid-1990s it was proposed that quantum effects in proteins known as microtubules play a role in the nature of consciousness. The theory was largely dismissed due to the fact that quantum effects were thought unlikely to occur in biological systems, which are warm and wet and subject to decoherence. However, the development of quantum biology now suggests otherwise. Quantum effects have been implicated in photosynthesis, a process fundamental to life on earth. They are also possibly at play in other biological processes such as avian migration and olfaction. The microtubule mechanism of quantum consciousness has been joined by other theories of quantum cognition. It has been proposed that general anaesthetic, which switches off consciousness, does this through quantum means, measured by changes in electron spin. The tunnelling hypothesis developed in the context of olfaction has been applied to the action of neurotransmitters. A recent theory outlines how quantum entanglement between phosphorus nuclei might influence the firing of neurons. These, and other theories, have contributed to a growing field of research that investigates whether quantum effects might contribute to neural processing. This review aims to investigate the current state of this research and how fully the theory is supported by convincing experimental evidence. It also aims to clarify the biological sites of these proposed quantum effects and how progress made in the wider field of quantum biology might be relevant to the specific case of the brain.
q-bio.NC
q-bio
Quantum effects in the brain: A review Betony Adams1 and Francesco Petruccione1 Quantum Research Group, School of Chemistry and Physics, and National Institute for Theoretical Physics, University of KwaZulu-Natal, Durban, 4001, South Africa. (Dated: 21 October 2019) In the mid-1990s it was proposed that quantum effects in proteins known as micro- tubules play a role in the nature of consciousness. The theory was largely dismissed due to the fact that quantum effects were thought unlikely to occur in biological systems, which are warm and wet and subject to decoherence. However, the de- velopment of quantum biology now suggests otherwise. Quantum effects have been implicated in photosynthesis, a process fundamental to life on earth. They are also possibly at play in other biological processes such as avian migration and olfaction. The microtubule mechanism of quantum consciousness has been joined by other the- ories of quantum cognition. It has been proposed that general anaesthetic, which switches off consciousness, does this through quantum means, measured by changes in electron spin. The tunnelling hypothesis developed in the context of olfaction has been applied to the action of neurotransmitters. A recent theory outlines how quan- tum entanglement between phosphorus nuclei might influence the firing of neurons. These, and other theories, have contributed to a growing field of research that inves- tigates whether quantum effects might contribute to neural processing. This review aims to investigate the current state of this research and how fully the theory is supported by convincing experimental evidence. It also aims to clarify the biological sites of these proposed quantum effects and how progress made in the wider field of quantum biology might be relevant to the specific case of the brain. 9 1 0 2 t c O 7 1 ] . C N o i b - q [ 1 v 3 2 4 8 0 . 0 1 9 1 : v i X r a 1 I. INTRODUCTION The idea that quantum physics might have something to do with explaining that most mysterious organ, the brain, has generated scepticism among scientists. Just because quan- tum theory and consciousness are both complex concepts does not mean they necessarily inform each other. Indeed the contexts in which both occur have long been assumed to be incompatible. The brain is a biological system, functioning at physiological temperatures and subject to the myriad interactions by which living organisms survive. Quantum effects, on the other hand, are conventionally limited to low temperature systems isolated from the detrimental effects of environmental interaction. However, the mutual exclusion of biological and quantum systems is no longer so absolute. Research in the field of quantum biology has had some success in identifying how quantum processes might benefit living organisms1 -- 5. On a fundamental level, it can be said that all biological systems are quantum mechanical, being composed of atoms and thus subject to the quantum theory of atomic structure first developed by Bohr and Rutherford at the beginning of the twentieth century6,7. In the field of quantum biology these are considered to be trivial quantum effects, what is more interesting is whether quantum phenomena such as coherence, entanglement and tunnelling might play a non-trivial role in enhancing the efficacy of biological processes. There is mounting evidence that quantum coherence might contribute to the extraordinary efficiency of photosynthesis. The avian compass has also been suggested to exploit quantum effects with behavioural evidence supporting the hypothesis. Olfaction, enzyme catalysis and the intricacies of DNA have all fallen under the scrutiny of researchers working in the field of quantum biology1 -- 5. The basic biology of the brain, elevated though it is by the inexplicable phenomenon of consciousness, is perhaps not, on a mechanistic level, so very different from other processes that take place in the body. It is no wonder then that interest has grown in the detailed physiological mechanisms that constitute the central nervous system. Being that it is still not perfectly understood exactly how the brain and its related systems work, quantum theory might have at least some contribution to make. There have been attempts to tackle the hard problem of consciousness8,9, or apply quantum theory to human psychol- ogy and cognition10 -- 12. This review, however, focuses on the various instances in which it has been suspected that quantum effects play a role in the structural mechanisms by which the brain performs its integral functions: the firing of nerves; the actions of anaesthesia, 2 neurotransmitters and other drugs; the sensory interpretation and organised signalling that is central to the vast neural network that we identify as our self. A. The classical brain Neuroscience has made great strides in understanding how the brain works. However, if our model of the brain were already fully realised there would be no need to investigate whether quantum theory might offer any insights. The question of how the brain works is, to state the very obvious, complicated. On one level it is a matter of matter, the network of cells and signalling processes that constitute the central nervous and related systems. But there is also the question of how this physiology gives rise to the phenomenon of consciousness. While advances in imaging techniques have resolved some of the structure of the brain, new discoveries are still being made. Recently, previously unheard of lymphatic vessels were discovered in the meninges of the brain, suggesting a link between the central nervous system and the immune system and prompting research into their role in neurodegeneration13,14. Even less well understood is the relationship between structure and function and to what extent functional connectivity, or an understanding of the mind, can be explained by the basic anatomy of the brain15 -- 18. 1. Brain organisation The central nervous system is made up of the brain and the spinal cord. The human brain is a complex organisation of neural tissue, with approximately 86 billion neurons19. While neurons are responsible for the electrical activity of the brain they are supported by glial cells, which have a number of functions20. Brain matter consists of both grey and white matter, where the former is mainly cell bodies and the latter is predominantly made up of myelinated axons that allow for the transport of signals and the connectivity of the different brain sections21. The brain is a network, dependent on the complex interaction of its different constituent parts, thus the assignment of specific function to specific region is a simplification. Nevertheless, for classification purposes it is often divided and subdivided into a number of regions, the simplest of these divisions being the fore, mid and hindbrain. 3 FIG. 1. The basic structure of a nerve cell. Although a number of the organelles are missing this simplified schematic is sufficient to gain some understanding of the location of the various quantum effects discussed in this review. Of interest are the microtubules which occur in different arrangements throughout the nerve cell and the mitochondria, which will be discussed in the context of electron transfer. The myelin insulated axon has also been proposed as a wave guide for neural photonics. The meeting of dendritic spine and axon terminal at the synaptic cleft is enlarged for clearer understanding of the mechanism of neurotransmission and the transfer of neural signals. Synaptic vesicles are also the site of the proposed uptake of Posner molecules during endocytosis. 2. Neural action Nerve cells, the main constituents of the central nervous system, are elongated cells con- sisting of cell body, dendrites and axon. A complex network of neurons throughout the body allows for the propagation of information that is facilitated by the firing or not fir- ing of neurons, the measure of a neuron's action potential. The biophysical mechanism of action potential in neurons is conventionally understood through the work of Hodgkin and 4 Huxley22. In order for a neuron to fire the resting potential must be raised to the requisite threshold potential. This is mediated by the gradient of electrically charged ions distributed across the cell membrane. Communication between neurons is integral to their ability to convey information. There is still some contention as to exactly how nerves communicate, with evidence for both chemical and electrical signal transmission23. This review will focus on the former, where neural communication is mediated by the release of neurotransmitters. When an action potential propagating along a neuron reaches the axon terminal it triggers the opening of voltage gated ion channels which in turn stimulate exocytosis, or the release of neurotransmitters into the synaptic cleft24. These neurotransmitters diffuse across the synaptic cleft and bind to special receptors on the dendritic spines, opening other ion chan- nels. Ions can now enter this nerve cell and change the membrane potential, potentially generating an action potential in the post-synaptic nerve cell24. While this is a simplistic model of the action of neurons it serves as a means to locate the quantum effects discussed in the next section. Although neurons have a number of constituent parts, those important in the context of this review are as follows. Microtubules, as the location of effects described in theories of consciousness as well as coherent quantum transport. Mitochondria, as the site of electron transport processes. The axon, as mediator of electrical signals and possible site of biophoton transfer. A basic understanding of the synaptic mechanism will also be beneficial to the discussion of both quantum effects in neurotransmission as well as the neural action of Posner molecules. 3. Consciousness While this review is preoccupied with the simpler question of structure, that is, the quantum physics of certain physiological mechanisms in the brain, it does include some theories as to how this physiology manifests as consciousness. Biology based attempts to explain consciousness include identifying its neural correlates by means of neuroimaging methods which study the changes in neural activity between conscious and unconscious states as well as altered states of consciousness25 -- 30. A formal description of consciousness, given the difficulty of quantifying its subjective experience, would likely borrow from complex network theory as well as disciplines from physics and philosophy15,31 -- 35. The question is still open as to whether quantum physics has something to add to the debate. 5 FIG. 2. The non trivial quantum effects discussed in this review and their neural context. B. Non trivial quantum effects In 1913, Bohr presented his model of the atom where, in contrast to classical orbits, electrons occupied discrete energy levels. This followed quickly on the heels of Planck's ad- vances in understanding blackbody radiation and Einstein's explanation of the photoelectric effect which, along with Compton's work with X-rays, ushered in the new era of quantum mechanics6,7. Fundamentally, all biology can be described as being quantum mechanical in the same way that all matter is quantum mechanical. However, the aim of this review is to investigate non trivial quantum effects in biological systems, to widen the scope of quantum theory to include biological mechanisms and not merely the description of their constituent atoms. Quantum weirdness is a well documented phenomenon. First Planck and Einstein demonstrated that radiation, normally understood as behaving like a wave, can also behave like a particle. De Broglie then suggested that matter, which seems dis- crete, can sometimes show wave-like effects such as interference. Despite being famous for its uncertainty, the formalisation of the theory has proved extremely successful in describing the behaviour of microscopic systems. The mathematical framework of quantum mechanics associates a physical system with a quantum state that contains all possible information on the system. What is interesting is that within this framework, for two quantum states describing a system, a linear combination of these states also describes the system. It is this that gives rise to the uniquely quantum effect of superposition states, one of the non- trivial effects discussed in this review. The concept of quantum coherence, which quantifies 6 the relationship between states in a superposition, is investigated in various biological con- texts. As is entanglement, the non-classical correlation between different quantum states. Entanglement is often discussed with reference to quantum spin, the property of elementary particles that determines their behaviour in a magnetic field. Electron tunnelling is also raised as a potential candidate for quantum effects in biological systems. An explanation of tunnelling follows from the probabilistic description of quantum mechanics which allows the possibility of a quantum particle passing through a classically forbidden potential barrier6,7. This review is particularly focused on inelastic tunnelling, in which the tunnelling electrons are coupled to vibrational modes in their biological context. II. QUANTUM EFFECTS IN NEURAL PROCESSES It is perhaps misleading to talk about quantum processes in the brain. Nerve cells extend throughout the body. The quantum processes discussed in this review are not confined to the brain but take place within neurons and at synapses. They are therefore implicated in the biological functioning of the entire body. It might be more accurate to describe these effects as quantum enhanced neural processing. A. Orchestrated Objective Reduction Orchestrated objective reduction (Orch OR), the application of quantum mechanical for- malism to the question of consciousness, was proposed by Stuart Hameroff and Roger Pen- rose in the 1990s9,36,37. In his 1989 book The Emperor's New Mind: Concerning Computers, Minds and The Laws of Physics Penrose addresses the possibility that the laws of classical physics are not sufficient to explain the phenomenon of consciousness, suggesting instead that quantum physics might be integral to this explanation8. His hypothesis initially lacked a biological context in which these quantum effects might occur. Hameroff, an anaesthesi- ologist by training, had been previously interested in microtubules and suggested that they could be a contender in which to situate a quantum model for consciousness. Hameroff and Penrose have subsequently collaborated on developing and refining the theory of Orch OR, by which quantum computations in microtubules influence neural firing and by extension constitute the neural manifestation of consciousness9. Microtubules in general have elicited 7 interest in various researchers attempting to model quantum effects in the brain. 1. Microtubules Microtubules are formed by the polymerisation of tubulin dimers, which consist of α and β tubulin proteins. Tubulin dimers first form longitudinal protofilaments; 13 of these protofil- aments then form a microtubule with diameter of approximately 25 nm9. Microtubules form part of the cytoskeleton of eukaryotic and some prokaryotic cells and contribute to cellular shape and structure. They have a variety of functions. They are integral to cell division, forming the spindle apparatus that mediates the division of chromosomes into daughter cells. Microtubules also act as tracks along which motor proteins move cellular constituents within the cell9,38,39. Whereas microtubules are present in all eukaryotic cells the theory of Orch OR is focused on microtubules in nerve cells and in particular those found in the dendrites and cell body (soma) of these cells. This is because microtubules in axons and non-neural cells have a radial, regular arrangement that is arguably less supportive of in- formation processing. Microtubules in the dendrites and soma are less regularly arrayed, forming what Hameroff and Penrose refer to as recursive networks well suited to learning9,40. Microtubules in non-neural cells are also dynamically unstable, able to disassemble in var- ious ways. Microtubules in dendrites and nerve cell bodies are prevented from disassembly by microtubule associated proteins, rendering them more stable and able to encode the long-term information necessary to the theory of Orch OR9,41,42. The specific composition of tubulin has also lent strength to quantum models of neural processing due to the fact that it is partially composed of chromophores such as tryptophan, arranged in a manner similar to photosynthetic systems in plants and bacteria43,44, which have been surmised to support coherent quantum effects. 2. The quantum model of Orch OR Hameroff and Penrose hypothesise that quantum computations encode information in microtubules and objective reduction is how this quantum information results in a classical output. The details by which Penrose's conception of quantum gravity gives rise to the objective reduction of the quantum wavefunction are beyond the scope of this review. In 8 this case Hameroff and Penrose's 2014 review of the theory is instructive9. The choice of microtubules, and more specifically tubulin dimers, as the biological context in which Orch OR takes place has been motivated by various reasons. It has been suggested that the manifestation of consciousness is not axonal firing but rather the signal integration that occurs in the dendrites and cell bodies of nerve cells. This is given some support by the fact that gamma wave synchrony, which has been suggested to be the neural correlate of consciousness, is generated by dendritic-somatic integration potentials9,45. As outlined above, microtubule arrangement in the dendrites and cell bodies of nerve cells is suit- able for information processing, making them a good contender for the biological site of consciousness9,40,42. The localisation of Orch OR in microtubules also offers a way to model a biological qubit, which is integral to the quantum nature of the theory. A qubit, the basic unit of quantum information, is a two-state system that can exist in a superposition of both states at the same time. Initially Hameroff and Penrose proposed that tubulin dimers might exist in a superposition of mechanical conformations coupled to London force dipoles9,37,46. More recent iterations of the theory locate the quantum description firmly in the constituent aromatic rings (phenylalanine, tyrosine and tryptophan) that make up the tubulin proteins. These have pi orbital electron clouds that demonstrate spatial delocalisation, giving rise to London force electric dipoles, that can exist in superposition. While Orch OR was originally formulated with electric dipoles in mind, the authors now propose magnetic dipoles related to electron spin9. 3. Discussions around Orch OR This review gives only a basic overview of the theory of Orch OR as a means to introduce the idea of quantum models of the brain and their biological sites of action. There have been a number of objections raised against Orch OR, the details of which are given in detail in Hameroff and Penrose's 2014 review9. Those responses that deal with the structural viability of quantum effects in the brain are briefly addressed in this section. The problem of decoherence is an issue commonly cited against the application of quantum theory to biological systems. Quantum mechanics is conventionally applied to isolated systems at low 9 temperatures whereas biological systems are typically described as being warm, wet and messy. In 2000 Max Tegmark addressed the question of decoherence in the context of Orch OR, concluding that calculated decoherence time scales of 10-13 -- 10-20 seconds are much too short for quantum effects to play any role in cognitive processes47. Hagan et al. responded to this by asserting that Tegmark had based his calculations on a model that did not closely resemble the one they had proposed. After recalculation for a more accurate model they conclude that decoherence times are closer to 10-5 -- 10-4 seconds48. With the development of quantum biology the question of whether quantum mechanics can contribute any meaningful insight into the functioning of biological organisms has shifted towards a positive answer. In particular the study of photosynthesis has revealed that the efficacy of energy transfer and charge separation might be enhanced by quantum effects49 -- 54. Thus the decoherence argument against Orch OR perhaps holds less weight than other evidence contradicting the theory. There has been some argument over the lattice arrangement of tubulin dimers in mi- crotubules, which can be of two types: A and B. The predictions of Orch OR favour type A whereas actual mouse brain tissue points to the predominance of type B. Proponents of Orch OR argue that this does not necessarily discount the theory as quantum effects may only occur in that fraction of microtubules that are geometrically suitable9,55. Objections by McKemmish et al. include the proposed conformational switching of the tubulin pro- teins that constitute the biological qubit. They note that the theory demands significant changes in tubulin structure and that any processes that might drive these conformational changes would be metabolically expensive56. More recent reformulations of Orch OR do not require such extreme conformational switching and the different states necessary for the qubit can be achieved through superposed electric or magnetic dipoles in aromatic rings9. McKemmish et al. also note that the electrons in a single aromatic ring cannot exist in superposition states, being completely delocalised. This can be incorporated by considering the electron clouds of two or more rings9. Reimers et al. take issue with the possibility of strong Frohlich condensation in microtubules57. They explain megahertz coherence demon- strated in microtubules by Pokorn´y58 as being due to weak classical Frohlich condensation. They go on to suggest that as Orch OR involves coherent strong Frohlich condensation the theory is flawed57. Hameroff and Penrose respond by citing experimental evidence of 10 the discovery of gigahertz, megahertz and kilohertz resonance in single microtubules. They argue that the experimental evidence argues strongly in the favour of Orch OR9,59,60. Other criticisms have been aimed at the objective reduction part of the theory, and the way in which this translates into consciousness. These have been addressed in detail by Hameroff and Penrose9. 4. Microtubules as a site for quantum effects While Orch OR theory is not without controversy, the suitability of microtubules as a site for quantum effects has given rise to a number of related approaches. It seems probable that neuronal microtubules are indeed implicated in consciousness and cognition. This is supported by the fact that certain chemicals that influence both consciousness as well as cog- nitive function, such as general anaesthetics and antidepressants, involve microtubules9,61,62. The proposed quantum action and alternative biological sites of processes related to both anaesthetics and antidepressants will be discussed in the sections B and C. Craddock et al. suggest an additional way in which microtubules might play a quantum role in neural processing. They take, as a comparative example, evidence of quantum beats in the light- harvesting complexes of plants and bacteria. They then suggest that the tryptophan residues present in the tubulin proteins that constitute microtubules are structurally and functionally capable of supporting the possibility of quantum coherent energy transfer43. The details of this research will be discussed further in the context of electron transfer processes in section D. B. Quantum models of general anaesthesia One of the ways in which we might understand the mechanism of consciousness or, less ambitiously, the details of neural processes, is by looking at chemicals that disrupt these processes. As Luca Turin puts it, 'the only thing we are sure about consciousness, is that it is soluble in chloroform'63. To this end the study of general anaesthetics has offered some way of structuring research into quantum neural processing. There are a few theories 11 as to how quantum effects are implicated in the action of general anaesthetics. One of these follows from the investigation of possible quantum effects in microtubules, initiated by Hameroff and Penrose's Orch OR theory64. Another looks at spin changes in anaesthetised fruit flies65. Both of these are preoccupied with the possibility that anaesthetic action disrupts electronic activity, a theory that first made its appearance in the 1980s66 -- 69. There is also some suggestion that the nuclear spin of anaesthetic molecules might influence their efficacy70. 1. The action of anaesthetics in tubulin proteins Craddock et al. argue that understanding general anaesthetics as having a network71 or synaptic based effect72 does not explain how anaesthetics inhibit the cognitive abilities of sim- ple single celled organisms such as slime moulds73,74. The fact that this cognition is linked to cytoskeletal microtubules points to a possible site for the action of general anaesthetics64,75. There is some uncertainty concerning the exact biological location targeted by anaesthetics. The Meyer-Overton rule links the potency of anaesthetics to their increased lipid solubility. Further research suggests that anaesthetics act in lipid-like hydrophobic regions in proteins64. Anaesthetics bind to a number of membrane and cytoplasmic proteins76 -- 78, Craddock et al. propose that tubulin, the protein subunit of microtubules, seems the most likely due to the fact that gene expression after exposure to anaesthetic compounds is concentrated on mi- crotubule dependent functions64,79. The quantum mechanism by which anaesthetics operate in microtubules is based on a number of theoretical observations. Building on research that tryptophan rings arranged favourably in tubulin proteins can act as quantum channels sup- porting coherent electron dynamics in a manner similar to photosynthesis43,80,81, Craddock et al. argue that anaesthetic gas molecules binding in these channels inhibit quantum effects and disrupt coherent energy transfer and that this is responsible for the effects of general anaesthetics on consciousness64,75. In a later paper Craddock et al. expand their theory to the prediction of anaesthetic potency. They argue that anaesthetic and related gases change collective terahertz dipole oscillations in a way that predicts the potency of their action82. 12 2. Quantum spin and general anaesthetics In his early research relating to quantum biology, Turin hypothesised that the sense of smell, understood classically as a lock-and-key mechanism, might instead be better ex- plained by quantum theory. More recently he has addressed the common thread between the very different molecules that act as general anaesthetics. Turin's theory is that anaesthetic molecules perturb electron currents in their target proteins65. The molecules that comprise the group of general anaesthetics range from the structurally simple noble gas xenon to the much more complex molecule alfaxalone, a range that includes numerous other chemicals without any apparent similarities. It is this structure-function anomaly that is motivation to look at the underlying physics of anaesthetics65,83. Being that it is structurally the simplest of the anaesthetics, more than one attempt has been made to understand the anaesthetic action of xenon. In a recent paper Li et al. examine the differing anaesthetic effects of xenon isotopes. Xenon has nine stable isotopes. Seven of these have zero nuclear spin but xenon 129 has a nuclear spin of 1 2 and xenon 131 of 3 2 70. In their experiment Li et al. compare the loss of righting reflex in mice, which correlates with consciousness, under the influence of the differ- ent isotopes. In order to exclude electronic effects they also calculate the polarisabilities of the different isotopes, finding them to be undifferentiated. Their results demonstrate that xenon isotopes with non-zero nuclear spin have a lower anaesthetic effect. As the results do not show a correlation between anaesthetic effect and atomic mass they conclude that the differences must depend on the value of the nuclear spin. Although the details of the mechanism are not clear the authors hypothesise that as spin half particles have been shown to be better suited to quantum entanglement, perhaps entanglement promotes consciousness in opposition to the effects of anaesthetics70. Their results recall research done by Fisher et al. into a possible mechanism by which neural entanglement might proceed84 -- 86. Of particular interest here perhaps are the isotope-dependent effects of lithium, another drug that alters conscious experience. In order to investigate whether quantum physics plays any common role in their mech- anism of action Turin et al. also focus on the physics of the xenon atom. In 1989 IBM 13 used a scanning tunnelling microscope to manipulate 35 xenon atoms on a nickel surface into spelling out the IBM logo; where the xenon atoms extend the conducting surface of the nickel87. In a similar manner, Turin et al. suggest that xenon atoms extend the highest occupied molecular orbit (HOMO) of the protein they interact with, facilitating electron transfer, and that this effect is common to the other molecules that act as gen- eral anaesthetics65. They test this hypothesis with an electron spin resonance experiment using anaesthetised fruit flies, looking for increases in spin caused by general anaesthetic. Discounting free electron spin signals from melanin pigments the results of the experiment lead them to conclude that general anaesthetics cause a change in spin. Further verification for the hypothesis is offered by the fact that spin changes differ for anaesthetic resistant flies. In addition to this, using density functional theory they demonstrate that chemicals which display similar but less exaggerated central nervous system effects to anaesthetics, also extend the highest occupied orbital momentum but to a lesser degree65. While electron spin changes under the influence of general anaesthetics points towards a possible quantum mechanism, there is still some scepticism as to whether the experimen- tal evidence is sufficient proof of principle. Turin et al. themselves note that the spin changes could be due to melanin. Although they checked their results against flies deficient in one of three possible melanins, they concede that particularly neuromelanin, the grey in grey matter, might play a role65. This is potentially interesting in the context of recent research investigating the quantum role of neuromelanin facilitated electron transport in sections of the brain88. Turin et al. also document the fact that the experiment was performed under distinctly unphysiologic conditions. In order for the flies to remain immobile and not disturb the readings they were kept at 6 C. Although some temperature variation was tested this was only between 2 and 10 C. In addition to this the flies were kept in anoxic conditions to isolate the effects of anaesthetic gases from those of oxygen. The potential role of oxygen in the experimental results points to the involvement of respiration, more specifically the movement of electrons in the electron transport chain of mitochondria89. Turin et al. cite a number of instances in which it has been shown that mitochondria are involved in the action of anaesthetics65,90 -- 92. In a more recent collaboration with Turin, Gaitanidis et al. invite speculation on an unpublished preliminary report of spontaneous radiofrequency emission from fruits flies subjected to a magnetic field. They suggest that 14 these emissions originate from the nervous system due to the fact that they stop under the influence of chloroform anaesthetic and that they are related to spin-polarised electron currents in cells. The authors note that these cellular currents could reflect mitochondrial metabolism but that the variable signal and reaction to anaesthetic suggests some more complex biological activity connected to neuronal activity93. 3. The outlook on quantum models of general anaesthesia While the comparison of electron dynamics in photosynthetic and microtubule proteins is largely theoretical the authors suggest its experimental verification is possible via two- dimensional electronic spectroscopy, a technique well established in the analysis of photo- synthetic systems5. They propose that quantum beating in tubulin will be altered under the influence of anaesthetics and that this is a measurable phenomenon64. While this has yet to be attempted there has been some recent experimental evidence that supports the possi- bility that quantum effects play a role in the action of anaesthetics. Using various types of spectroscopy, Burdick et al. report that, in contrast to nonhalogenated ethers, halogenated ethers interact with entangled photons of specific wavelength94. C. Neurochemical binding and activation mechanisms For the purposes of this review the term neurotransmitter is used with respect to a variety of neurochemicals that fall into this category, binding to neuroreceptors and modu- lating neural signalling. Neurotransmitters emitted by one nerve cell bind to receptors on an adjacent nerve cell and facilitate the opening of ion channels. This is fundamental to the generation of action potentials and disruption of this process is believed to contribute to mental illnesses95,96. Although this review focuses on the action of neurotransmitters it has been suggested by Vaziri et al. that the mechanism by which ion channels allow the selective transmission of ions might also not be strictly classical, showing evidence of quan- tum coherence97. The binding action of neurotransmitters is conventionally understood as a lock and key mechanism whereby the shape of the neurotransmitter matches its specific receptor98,99. The lock and key or docking mechanism is implicated in a number of biological processes: neurotransmission; the action of enzymes; olfaction; DNA binding, all depend to 15 some extent on lock and key theory98,100 -- 102. Despite the success of the theory however, an alternative view suggests that something more than molecular shape might be necessary to explain olfaction and, more recently, neurotransmission. 1. Lock and key or quantum vibration Neurotransmitters are a class of molecules that bind to G protein-coupled receptors (GPCRs). GPCRs are a group of receptors that, upon detection of appropriate molecules known as ligands, activate signalling pathways in eukaryotic cells. GPCRs play an impor- tant role in medical innovation, as targets for drug action103. In addition to neurotrans- mitters, there are a number of biological molecules that bind to GPCRs; these include odourants104. Olfaction is classically understood as depending on the respective shapes of the chemical/receptor pair. However, an alternative vibrational theory of olfaction was first developed by Dyson as long ago as the 1930s105,106. In his 1996 paper Turin suggested that the vibrational frequency of a given odourant contributes to quantum tunnelling at the receptor107. In 2011 Franco et al. presented experimental evidence in support of the theory. In particular they claimed that fruit flies could differentiate between deuterated odourants108. There is also some evidence that more complex species such as lake white- fish and the American cockroach can differentiate between isotopes of amino acids and pheromones109 -- 111. Horsfield et al. document a number of experiments that test the the- ory on both the behavioural and physiological level105. Deuterated odourants are a useful means to test the vibrational theory of receptor activation due to the fact that replacing hydrogen by deuterium adds vibrational modes, for instance the 2150 cm−1 vibration of the carbon-deuterium bond105,112,113. However, there has also been some recent discussion as to the fact that a differential olfactory receptor response between undeuterated and deuterated odourants could in fact be due to a minute contaminant in one of the samples. Paoli et al. conclude that although their results do not prove the vibrational theory of olfaction they also do not disprove it, merely calling for caution in experimental approach113. While there is little verification for a vibrational olfactory mechanism in mammals, a number of the experiments support the theory for the case of insects. It should be noted, in a discussion of GPCR mechanisms, that insect olfactory receptors differ from mammalian G- protein related olfactory receptors and thus experimental results may not be generalisable105. 16 While the theory remains controversial it has recently been re-examined in the context of neurotransmission. A number of different neuroreceptors have been investigated. Adenosine receptors are rhodopsin-related G-protein coupled receptors that bind the neuromodulator adenosine, but can also bind a variety of other agonist and antagonist molecules114. The stimulating action of caffeine, for instance, is due to its antagonistic binding to adenosine receptors. Because GPCRs are such an important target for pharmaceutical intervention the classification of molecules that act as agonists or antagonists is a well-developed re- search field. Molecules can be classified using various molecular descriptors which include information from molecule structure, topology and geometry to dipole moment, electric polarisability and electrostatic potential114. Following on from the fact that adenosine re- ceptors and olfactory receptors are both class A GPCRs, Chee and Oh present research that tests whether vibrational frequency might also be an effective molecular descriptor. They suggest that classifying ligands by vibrational frequencies is an effective way of discrimi- nating agonist from antagonist114. Chee et al. then refine this research using a machine learning approach and conclude that selected features of molecular vibration allow for ligand classification of adenosine receptor agonism115. Whereas Chee et al 's research investigates adenosine receptors Hoehn et al. focus in partic- ular on the neurotransmitter serotonin and its receptors109,116. Although it has a number of other functions in biological systems serotonin is perhaps most well known for the role it plays in mood. A widely prescribed class of antidepressants, the SSRIs (selective serotonin reuptake inhibitors) target this neurotransmitter117. In their research Hoehn et al. address the possibility that serotonin neurotransmission might utilise vibration assisted inelastic tunnelling effects. Using inelastic electron tunneling spectroscopy (IETS) theory that was first developed to understand olfaction they investigate the tunnelling spectra of endogenous and non-endogenous agonists that bind to the serotonin receptor. Non-endogenous agonists of this receptor include LSD (lysergic acid dimethylamide), DOI (2,5-dimethoxy-4-Iodo- amphetamine) and other psychedelic phenethylamines109. Their results suggest that the serotonin molecule shares an inelastic electron tunnelling spectral peak with other agonists that activate the serotonin receptor. Although the lock and key mechanism has been very successful in modelling certain aspects of the action of signalling proteins one of the ways in 17 which it falls short is the prediction of agonist potency. In addition to the shared spectral peak of related agonists Hoehn et al. also report that the intensity of this peak might be used as a predictor of agonist potency109. 2. Experimental verification of the vibration assisted tunnelling hypothesis Experimental verification of the vibrational theory of neuroreception follows the lead of Franco et al. in the context of olfaction, where the effects of deuterated odourants are investigated108. Hoehn et al. report the results of an experiment to test the feasibility of their theoretical approach by measure of receptor affinity and activation. Once again, ago- nists of the serotonin receptor were chosen, specifically 2,5-dimethoxy-4-iodoamphetamine (DOI) and N,N-dimethyllysergamide (DAM-57). However the experiment failed to confirm the vibrational theory of neuroreception as they report that selective deuteration had no influence on binding affinity or activation116. In another study regarding the binding mechanism of the neurotransmitter histamine, authors Krzan et al. report a significant distinction in the binding patterns of histamine and its deuterated counterpart, as well as for other agonists of the histamine receptor118. In particular they found that deuterating histamine increased its binding affinity. They discuss this in the context of the vibrational theory of olfaction and GPCRs more broadly. Instead of confirming the theory they propose an alternative reason for the fact that ex- periments suggest that animals can differentiate between deuterated odourants, concluding that it is a nuclear quantum effect governed by differences in the strength of hydrogen bonds before and after deuteration118. In contrast to this, theoretical work comparing the structure-activity/vibration-activity relationship of the histamine receptor and its various ligands suggests that molecular vibration does play some role in ligand function. In the study 47 ligands that bind to histamine receptors were investigated using a computational approach to molecular vibration. This led to the conclusion that the many varying agonists and antagonists can be to some extent classified by their molecular vibrations119. 18 3. The functional mechanisms of GPCRs in general The lack of experimental verification for vibrational quantum effects in the context of central nervous system GPCRs prompts Hoehn et al. to suggest that either olfactory recep- tors function differently from other GPCRs or the vibrational theory of receptor activation is wrong. Gehrckens et al. address the latter criticism in a recent preprint, in which they use density functional theory to investigate the electronic structure of rhodopsin104. They choose to look at rhodopsin rather than olfactory GPCRs due to the fact that a high res- olution structure of rhodopsin is available for research purposes. Rhodopsin also shares important features with olfactory receptors and is considered the evolutionary ancestor of GPCRs120,121. The main aim of their research is to demonstrate a mechanism for electron transfer in olfactory receptors. They go on to identify a tryptophan donor and zinc accep- tor that would indeed allow electron transfer in rhodopsin, although they also emphasise the fact that this capacity for electron transfer does not necessarily play a role in the functionality of rhodopsin. In the context of whether this effect might be generalisable to GPCRs they conclude that an electron transfer mechanism might have been exploited by the offshoots of rhodopsin, in particular olfactory receptors. They do however suggest that in this sense olfactory receptors may be unique in that they are relatively non-specific, a requirement made necessary by the novelty and variety of odourants that will bind to them. Neurotransmission, they argue, depends on a binding that is much more specific. On the subject of neurotransmitters, they hypothesise that the anomalous binding structure of the antagonist metitepine to serotonin related receptors can be explained if the metitepine is in the form of a radical, having gained an electron in binding to the receptor. This mechanism is in line with the idea that GPCRs are electronic devices rather than simply lock-and-key104. Both theoretical and experimental results do not offer any clear conclusions with respect to the vibrational theory of neurotransmission for adenosine, serotonin and histamine recep- tors and their related ligands. Further research is necessary to clarify the viability of the approach and its specific relation to the different actions of neurotransmitter binding affinity and activation capability115. Understanding the mechanism or mechanisms of ligand-GPCR interaction is particularly important due to the fact that GPCRs are a major drug target associated with one third of all pharmaceuticals103,115. 19 D. Alternative signalling processes and biophotons It is perhaps interesting that GPCRs, in addition to binding with molecules such as neurotransmitters and odourants, also bind to photons122. There has recently been some suggestion that cellular communication and even possibly neural signalling might make use of biophotons in addition to more well established mechanisms such as neurotransmitters123 -- 125. Biophotons, are spontaneous ultra-weak photons in the near-UV to near-IR spectral range produced by biological systems, in particular through oxidative processes in mitochondria125,126. As noted in a recent paper on the subject, it is known that biophotons are produced in brains125. Correlations have been found between biophoton intensity and neural activity as well as oxidative dysfunction of neural cells in rat and mouse brains127 -- 129. There has even been some attempt to explain human intelligence in the context of biophoton emis- sion. Following on from a study that demonstrated that glutamate, a key neurotransmitter, can mediate biophoton production130, Wang et al. examine the spectral characteristics of biophoton emission in a range of species. They conclude that there is a correlation between higher order intelligence and the spectral redshift of biophotons emitted by sample brain slices. This redshift increases in the order of frog, mouse, chicken, pig, monkey, and human, which reflects with some accuracy the phylogenetic tree, although they concede that there is no completely objective means to measure intelligence across species131. Their argument has met with some scepticism, with Salari et al. responding that the experimental results insufficiently support the hypothesis and that without a mechanistic connection between spectral shift and intelligence, their correlation is more likely to be coincidence132. Biophotons have been implicated in cellular signalling in plants, bacteria and even kid- ney cells. Following from this Sun et al. present experimental evidence that nerve cells can also conduct biophotons, and that this effect can be inhibited by the application of a neural conduction block anaesthetic124,133. In their attempt to model the mechanistic details as to how biophotons are utilised in information transfer, Kumar et al. argue that neurons are well suited to photonics125,134. They cite evidence that nerve cells contain possible photon sources from mitochondrial respiration or lipid peroxidation, as well as photon detectors such as centrosomes and chromophores125,135 -- 138. They then hypothesise that myelin-coated axons of nerve cells act as waveguides for biophotons and that this might facilitate quantum effects 20 such as entanglement125. In order to demonstrate proof of principle the authors develop a theoretical model of light guidance in axons, solving the three dimensional electromagnetic field equations in the relevant context. They investigate a number of limiting factors and potential pitfalls and conclude that axons are mechanistically viable as waveguides. They go on to suggest various experiments in which the different aspects of their hypothesis might be put to the test125,134. 1. Quantum effects in electron transfer In a review that addresses the possibility of quantum neurobiology, Jedlicka suggests that cognition, typified by complex signal processing and integration, can potentially be thought of as occurring outside of strictly neural systems and that the quantum information processing abilities demonstrated by plants and bacteria could be useful to the study of cognition in higher animals10. Thus far, the theoretical and experimental advances made in understanding the role of quantum effects in biological systems have been primarily to do with photosynthesis, in particular light harvesting and energy/charge transfer processes5. Many of the theories of quantum effects in the context of neural processes focus on electron dynamics. As noted by Toole et al. coherent energy transfer in photosynthesis is less a unique feature of photosynthetic systems than it is the result of the specific arrangement of chromophores within a protein44. Kurian et al. propose that biophoton production in mitochondria is absorbed and channelled via resonant energy transfer by co-localised microtubules. They argue that neurodegener- ative diseases related to compromised microtubule networks could result from ineffective channelling of biophotons into signalling or dissipation. Experimental evidence shows that microtubules undergo organisational changes after exposure to photons, particularly in the absorption range of tryptophan and tyrosine139. As previously noted, a related paper presents a computational approach to quantum coherent energy transfer in microtubules43. The authors argue that similarly to the arrangements of chromophores in light harvesting photosynthetic complexes, the tubulin proteins that constitute microtubules have appropri- ate arrangements of chromophores such as tryptophan. They conclude that it is feasible that tubulin proteins could support coherent energy transfer in microtubules43,80. In a very 21 recent related paper Celardo et al. extend the similarities between photosynthetic antenna complexes and microtubules by demonstrating that tryptophans in microtubules can theo- retically exhibit superradiant excitonic states81. The possible role of tryptophans in electron transfer processes has also been previously addressed in this review with respect to the action of neurotransmitters in binding to and activating GPCRs. In their investigation of rhodopsin, Gehrckens et al. outline a mechanism by which a tryptophan donor and zinc acceptor facilitate electron transfer104. Tryptophan seems particularly interesting with respect to neural processes as it is the precursor to serotonin, an important neurotransmitter. This connection prompts Tonello et al. to hy- pothesise that the structure of consciousness emerges in a manner analogous to the serotonin dependent movement of plants towards their source of energy140. In neural processing this is extrapolated as biophoton harvesting by tryptophans and concomitant serotonin-mediated neural communication and plasticity140 In a later paper Tonello et al. expand on this idea with the hypothesis that the gastrointestinal-brain axis in higher animals might have evolved from the root-branch axis of plants, and that light plays an important role in both systems141. While these ideas remain for the moment largely theoretical, much of the theory of electron transport and resonant energy transfer has focused on the involvement of micro- tubules. Alternatively, the role of biophoton use and energy transfer in neural processes could be furthered by a closer look at the electron transport chain of mitochondrial respi- ration. As previously noted, mitochondria are already implicated in the action of general anaesthetics65,90 -- 92. They are also the primary production site of biophotons. They might serve as an alternative biological location in which quantum effects could be investigated. Research suggests that the arrangement of chromophores in bacteria allows for quantum coherence in their photosynthetic processes142. Similarities between the DNA of bacteria and eukaryotic mitochondria has led to the theory that mitochondria are descended from bacteria, though recent research suggests they are less directly related143 -- 146. While the ancestry of mitochondria remains uncertain their similarity to bacteria could arguably manifest in biological structures that support quantum effects. Both photosyn- thesis and mitochondrial respiration make use of an electron transport chain, which powers 22 FIG. 3. Simplistic comparison of the electron transport chains of respiration and photosynthe- sis. Electrons are transported through a series of complexes or photosystems, which results in the creation of a proton gradient. This is used to drive the production of ATP. While the specific complexes differ between chains they share some features relevant to the possibility of quantum processing. The presence of chromophores, for instance, which arguably support coherent energy transfer. While the interaction of light with the photosynthetic electron transport chain has been studied in detail, analogous effects in mitochondria are not well understood. Oxidative processes in the electron transport chain of mitochondria are thought to be the main source of biophotons. The function or target of these photons, however, is less clear. It has been proposed that microtubules close to mitochondria absorb and channel biophotons in a quantum coherent manner139. Mito- chondria themselves also demonstrate the capacity for photon absorption, in particular complex IV in the electron transport chain, which is the target of therapeutic application of infrared light to treat disorders of the brain151,153,158. 23 the charge gradient necessary for ATP synthesis. A group of researchers from the Quantum Biology and Computational Physics research group at the University of Southern Denmark are already investigating proton-coupled electron transfer in the cytochrome bc1 complex in photosynthetic bacteria and higher organism cellular respiratory systems. The group has published a number of papers motivated by the fact that malfunctions in the bc1 complex lead to many different diseases. They also hope to better understand photosyn- thesis in order to optimise energy conversion research147 -- 149. The reduction and oxidation of ubiquinone is essential to the electron transfer reactions in cytochrome bc1 and the concentration of ubiquinone is particularly high in the high energy consumption environ- ment of the brain150,151. There is some evidence that transcranial photobiomodulation, the application of red or near-infrared laser light to the cranial area, is an effective treatment for various brain disorders152,153. It has also shown promise in treating depression154 and when coupled with ubiquinone supplementation151. Research suggests that, among other effects, photobiomodulation can improve attention, memory, executive function and even rule-based learning155 -- 157. The mechanism of this effect is still not completely clear but the site of the effect is most likely the mitochondrial electron transfer chain, in particular the chromophores in the different complexes that constitute the chain151,153,158. There are a number of these chromophores, including tryptophan, which plays a central role in the various theories that espouse quantum effects in neural processing. Given that the role of quantum effects in light harvesting, electron and charge transfer has contributed to a better understanding of elements of the electron transfer chain in photosynthesis, this knowledge might be fruitfully applied to the mitochondrial electron transport chain. 2. Magnetic field effects and the brain The importance of tryptophans in energy transfer is evident in another well established field of quantum biology: the avian compass. A leading theory of bird migration suggests that birds utilise the earth's magnetic field as a guiding tool by means of the radical pair mechanism159 -- 161. The radical pair mechanism can be summarised in three main steps. First, a photon incident on a molecule causes electron transfer and pair formation. Second, the radical pair, originally in singlet spin state, interconverts between singlet and triplet state under the influence of the nuclear hyperfine and geomagnetic Zeeman effects. And 24 finally spin-dependent recombination leads to some signalling state that the bird interprets as a spatial directive161. It is generally accepted that the molecule in which this occurs is the blue light activated flavoprotein cryptochrome162 -- 165. Theoretical research, backed by spectroscopic evidence, suggests that light activated electron transport occurs between flavin adenine dinucleotide (FAD) and tryptophan residues161. Although it is well known that a number of species have a functioning magnetic sense, humans have not yet been added to that list. In a recent paper, Wang et al. present the results of an experiment that demonstrates the effects of earth strength magnetic fields on the human brain166. The authors report that magnetic field changes result in a decrease in amplitude of alpha frequency (8-13 Hz) brain waves, an effect normally associated with the brain's processing of external stimuli. They conclude that the effect is likely to be due to fer- romagnetism rather than the radical pair mechanism. This is because the effect is dependent on the polarity of the field, for subjects in the Northern Hemisphere the alpha wave response only occurs for horizontal rotations if the static component is directed upwards166. While this would appear to exclude the radical pair compass which is dependent on inclination, the corresponding effects would perhaps need to be confirmed for subjects in the Southern Hemisphere. It has been suggested, in the context of avian migration, that birds employ both ferromagnetism and the radical pair mechanism for navigation160,167,168 and this could also be true of humans' magnetic sense. The question being whether there is any evidence that might link magnetic effects in humans with the radical pair mechanism. Cryptochrome, the proposed site of the avian compass, is also present in humans169. Foley et al. use a transgenic approach to show that human cryptochrome can act as magnetosensor in the magnetoreception of fruit flies170. It is also potentially interesting that is it alpha waves that are effected by the magnetic field changes. Van Wijk et al. present evidence that biophoton fluctuation is correlated with the strength of alpha wave production, where they measure biophoton production in terms of the fluctuations of reactive oxygen species (ROS)124,171. The production of ROS and how this relates to the radical pair mechanism has been the motivation for various papers published in the field of quantum biology. Marais et al. outline a quantum protective mechanism in photosynthesis. They hypothesise that the high spin iron in the reaction centres of photosystem II of the electron transport chain 25 exerts a magnetic field effect that reduces the triplet yield of radical pairs that are formed during electron transfer. Triplet states are instrumental in forming ROS which are toxic to living cells172,173. In another paper, Usselman et al. show how yields of ROS in live cells are altered by radical pair dynamics, in particular coherent singlet-triplet mixing under the influence of oscillating magnetic fields at Zeeman resonance174. If ROS are correlated with alpha wave production124,171 and ROS yields are altered by the Zeeman effect174,175, then this might point to a mechanism that explains the influence of geomagnetic fields on alpha waves in the human brain. Reactive oxygen species participate in cellular signaling176 and have been implicated in aging and numerous diseases including mental conditions such as depression and schizophrenia177 -- 182. If humans do have physiological systems that depend on the dynamics of radical pairs then it is expedient to understand exactly how these function. The avian compass is disrupted by broadband radiofrequency electromagnetic radiation that is well below the WHO-recommended level183,184. This radiation might equally be upsetting the balance of ROS in other species and in turn leading to biological malfunction. There has been some suggestion that the radical pair mechanism is implicated in the development of can- cer through circadian rhythm disruption and related ROS levels185. Though the threat of radical pair mediated carcinogenesis is debatable, being equivalent, as one study finds, to the risk of travelling some kilometres towards or away from the earth's magnetic poles186. Nevertheless, the disruption of circadian rhythms has also been linked to mood disorders and cognitive function187 -- 189. Although the evidence remains contentious, there is also some documentation of the psychological effects of geomagnetic storms, which alter the earth's magnetic field for a limited period of time190 -- 192. The radical pair mechanism could offer a testable hypothesis as to how these effects occur190. E. Neural entanglement One of the first and strongest objections to quantum models of consciousness was the phenomenon of decoherence: that the non-ideal environment of biological systems would destroy any quantum effects before they could prove useful. As initially calculated by Max Tegmark in response to Orch OR theory, the timescales on which decoherence occurs in 26 FIG. 4. A scheme for neural entanglement. The Posner molecule structure is interpreted loosely here as a distorted cube with calcium ions (green) at each vertice and at the centre, and a phosphate ion (yellow) on each of the six faces198. This diagram is not meant to accurately convey the exact geometry of the Posner molecule but rather give some sense of Fisher's proposed qubit. The two Posner molecules share an entangled pair of phosphate ions. When two entangled Posner molecules are spatially separated and subsequently taken up into different presynaptic neurons in the process of vesicle endocytosis, their entanglement results in the correlated release of neurotransmitters and the coordination of firing across neurons84. the environment of the brain are considerably shorter than neural firing rates47. This con- clusion has subsequently been challenged not simply by researchers interested in Orch OR, but also by the development of the field of quantum biology. If quantum effects play a role in photosynthesis and possibly other biological contexts, then it is not such a stretch to consider that they may play a role in neural processes. Decoherence, it has been argued, 27 might even enhance energy transfer193. While research suggests that long-lived coherence in photosynthetic systems lasts for picoseconds194 and the lifetime of the radical pair mecha- nism is discussed in terms of milliseconds161, it seems unlikely that coherence in biological systems extends beyond these timescales. That is until a new hypothesis concerning neural entanglement proposed coherence that could last for hours or even days. 1. Entangled Posner molecules The hypothesis is this: that phosphorus nuclear spin might function as a neural qubit to allow for quantum processing to play a role in cognition84 -- 86. Following on from the proposal by Hu and Wu, that consciousness is linked to quantum spin195, Fisher argues that the spin-half phosphorus nucleus seems the only likely candidate for this role due to the fact that for any nucleus with spin greater than half, the electric quadropole moment means quicker decoherence due to electric field interactions in addition to the slower decoherence caused by the magnetic fields of nearby nuclei. Spin-half nuclei are thus more favourable in terms of decoherence times. Out of the various elements and ions that play an essential role in biological systems, only hydrogen and phosphorus have spin-half nuclei. Fisher goes on to describe the details of how the quantum information carried by phosphorus nuclei is created and preserved84 -- 86. Phosphorus is bound into phosphate or polyphosphate ions such as pyrophosphate. Phos- phate ions constitute part of the adenosine triphosphate (ATP) molecule, an organic chem- ical that acts as a source of energy for the many essential actions that sustain living organisms84,196. The electron transport chain in both photosynthesis and respiration cre- ates a proton gradient in order to drive ATP production. Fisher postulates that when adenosine triphosphate is hydrolysed to adenosine monophosphate and pyrophosphate the two phosphorus nuclei in the pyrophosphate ion will have a specific spin alignment, either a singlet or one of three triplet states. A quantum mechanical treatment of the enzyme catalysed reaction that creates two phosphate ions out of pyrophosphate demonstrates that this reaction depends on the nuclear spin state. More specifically, the enzyme conditional outcome of the reaction, where the spin dynamics of the triplet states are unfavourable, mean that the phosphorus nuclear spins in the two distinct phosphate ions will emerge in a 28 singlet entangled state84. 2. Coherence timescales for Posner molecules Phosphorus in the phosphate ion is surrounded by an oxygen cage, where the oxygen nuclei all have zero spin. Despite this the coherence time is still very short for solvated phosphates, approximately one second, due to the fact that hydrogen quickly binds to phos- phate and the non-zero proton spin contributes to decoherence. A coherence lifetime of a second is long enough for the quantum effects to be sustained over the process of cellular diffusion84. However in order for effects such as memory storage and retrieval to utilise quantum effects, coherence lifetimes need to be significantly longer. Fisher argues that if binding with the hydrogen can be pre-empted by a spin zero cation such as calcium, then longer coherence times might be possible. He identifies a possible molecule as the Posner molecule197, thought to be involved in the formation of hydroxyapatite, an important con- stituent of bone tissue198 -- 200. A Posner molecule is, simplistically, a distorted cube with a calcium ion at each vertice, a ninth at the centre and a phosphate ion on each of the six faces of the cube198. The entangled phosphate spins then result in entangled Posner molecules84. Fisher estimates that for phosphorus spins in Posner molecules coherence times could be as long as hours84 or, in a later paper, 21 days85. For solvated Posner molecules the mag- netic fields from protons in surrounding water molecules cause decoherence but, as Fisher argues, the rapid tumbling of Posner molecules means that the average magnetic field felt by a phosphorus nucleus will be zero. Decoherence will then only happen from residual magnetic field fluctuations and coherence times will thus be much longer84,85. Player et al. respond to this with their own calculation of coherence times for Posner molecules. They acknowledge that long lived spin states in nuclei are well accepted, with relaxation times of up to an hour being recorded198,201. However, they argue that these extended times are partly due to experimental control of the coherent spin dynamics, a condition that Fisher does not take into account198. In their analysis of Posner spin dynamics they take a number of other factors into account such as dipolar and scalar couplings within Posner molecules as well as the Zeeman interaction of the phosphorus spins with the geomagnetic field198. With these as the dominant relaxation pathway they arrive at a relaxation time of 37 minutes, 29 as opposed to Fisher's shortest estimation of approximately a day. They also discuss other ways in which this relaxation could take place even more quickly198. 3. Posner molecules in their biological context Fisher also outlines the way in which entangled Posner molecules give rise to quantum effects in neural processes. The entanglement of the spin and rotational states of Pos- ner molecules leads to correlations in the chemical reactions of spatially separated Posner molecules. These entangled molecules are taken into presynaptic glutamatergic neurons in the process of vesicle endocytosis. The acidic environment causes Posner molecules to bind and release calcium which stimulates exocytosis and further release of glutamate, which enhances neural firing. Thus, because the chemical reactions of Posner molecules are entan- gled the subsequent release of glutamate and resultant neural firing might also be considered entangled84,86. Fisher outlines a number of possible experiments that might verify the var- ious stages of the theory, most pressingly whether Posner molecules are present in bodily fluids. Whereas free-floating calcium phosphate clusters resembling Posner molecules have been found to be stable in simulated body fluids202, there is still some uncertainty as to whether they are present in vivo84,85. Less fundamental but more interesting in the context of this review is the suggested investigation of the effects of a chemically viable replacement of the central calcium in a Posner molecule with lithium ions84,85. There is some evidence that lithium, used to alleviate the symptoms of bipolar disorder, has isotope dependent effects on the behaviour of rats84,203. Fisher postulates that the efficacy of lithium as a treatment for mental disorders could be due to the increased decoherence induced by the lithium nuclear spins included in the Posner molecule84,86. 4. Neural qubits and quantum computing While the results of experiments to verify this model of entangled neural processes are yet to be completed the theory has caught the interest of researchers working in the field of quantum information theory. Halpern and Crosson apply Fisher's Posner molecule theory 30 in the context of quantum communication, quantum computation and quantum error cor- rection. They address how a quantum information based model of Posner molecules might contribute to the understanding of Posner chemistry and vice versa, what insights Posner molecules have for quantum information processing. They also demonstrate how entan- glement can change molecular binding rates, which goes some way to supporting Fisher's neural entanglement hypothesis204. The authors conclude that the quantum information based formulation of Posner molecules might be a way of framing biological Bell tests204, a claim that is potentially interesting in light of recent observations of entanglement between living bacteria and quantised light205. This is not the first time that phosphorus spin and quantum computing have crossed paths. In 1998 Kane proposed a scalable quantum computer where information is en- coded onto spin half phosphorus nuclei in a substrate of spin zero silicon, ensuring the long decoherence time necessary for quantum computing. Computations are then performed through the interaction of nuclear spin with donor electrons206. Kane's idea has recently been reworked by Tosi et al. into what they call the flip-flop qubit, where the combined electron-nuclear spin states of a phosphorus donor are controlled by microwave electric fields. This research, along with work done by He et al. into engineering a viable exchange interaction between two phosphorus bound electrons has done much to advance the course of spin based quantum computing207,208. The fields of quantum computing and quantum neurobiology might inform each other in other ways too. Quantum dots have been proposed as an alternative means to implement a quantum computer209. They have also been used to model the mechanism by which anti-acetylcholine receptor antibodies contribute to the neuromuscular disorder myasthenia gravis210. Even more recently, graphene quantum dots have been shown to prevent and even undo the protein clumping of neurons in the brain that leads to Parkinson's disease211. A separate study study demonstrated similar results for Alzheimer's disease212. III. CONCLUSION The field of neurophysics already attests to the fact that the study of the brain borrows from ideas across the spectrum of physics: electricity and magnetism, mechanics, thermo- 31 dynamics, optics15. It is perhaps not completely surprising that quantum physics might contribute too. While this review outlines working theories as to how quantum effects might be implicated in neural processes, the research remains largely theoretical. Although some experimental evidence points to the validity of certain of these theories, conflicting results mean it is difficult to draw any strong conclusions. However, many of the authors cited in this review suggest ways in which their theories might be put to the test experimentally. As already suggested by Craddock et al. advances in experimental techniques such as two- dimensional electronic spectroscopy, which has been successfully applied to understanding the quantum nature of photosynthesis, might yield similar success in the context of energy transfer processes that impact neural signalling64. A number of the hypotheses included in this review rework, to some extent, ideas already established with respect to other biolog- ical systems. Coherent energy transfer in photosynthesis is reimagined in the tryptophan rings of neural microtubules. The vibrational theory of olfaction is reapplied to the binding action of neurotransmitters. And while Fisher's Posner qubits depend on nuclear spin, un- like the radical pair model developed with respect to avian migration, their entangled spin dynamics mediate the outcome of signalling processes in a potentially analogous fashion. As it stands, the future of quantum neurobiology is the future of quantum biology more generally; progress made in either will further the other. The questions are then techni- cal. How exactly to apply experimental techniques to the specific neural context. How to refine, for example, experiments that measure the differing effects of isotopes without the confounding influence of contaminants. This mechanistic focus would seem to ignore the question of consciousness and how it emerges from these underlying neural mechanisms. Lynn et al. describe the future of brain network research, albeit in a classical capacity, as a cross-scale approach that links the microscopic to the macroscopic: protein reactions in neurons to synaptic connectivity to brain region connectivity to social networks15. Jedlicka, in his review of the future of quantum neurobiology, touches on the evidence that quantum physics has also been used to describe human behaviour10 -- 12. The use of a quantum frame- work to accurately describe human behaviour does not necessarily mean that this behaviour results from quantum effects in neural processes. A clearer understanding of these processes, however, could also elucidate to what extent quantum physics is involved in the emergence of cognition from neural activity10. Should it turn out that experimental evidence supports hypotheses of quantum enhanced neural processing we might turn this knowledge towards 32 thinking about how we think. FUNDING AND ACKNOWLEDGMENTS B.A and F.P. were supported by the South African Research Chair Initiative of the Department of Science and Technology and the National Research Foundation. B.A. was also supported by the National Institute for Theoretical Physics. Thank you to Angela Illing for the diagrams. REFERENCES 1J. Al-Khalili, J. McFadden, Life on the edge: the coming of age of quantum biology, London, UK Bantam Press (2014). 2J. McFadden and J. Al-Khalili, Proc. R. Soc. A 474, (2018). 3M. Mohseni,Y. Omar, G.S. Engel, M.B. Plenio (eds). Quantum effects in biology, Cam- bridge, UK Cambridge University Press (2013). 4N. Lambert, Y.N. Chen, Y. Cheng, C. Li, G. Chen and F. Nori, Nature Physics 9, 10 (2013). 5A. Marais et al., J. R. Soc. Interface 15, 20180640 (2018). 6Zettili N., Quantum Mechanics Concepts and Applications: Second Edition John Wiley & Sons Ltd, UK (2009). 7Haken H & Wolf HC., Atomic and Quantum Physics Springer-Verlag, Berlin Heidelberg (1987). 8R. Penrose, The emperor's new mind: Concerning computers, minds, and the laws of physics, New York, US: Oxford University Press, (1989). 9S.R. Hameroff and R. Penrose, Physics of Life Reviews 11(1), 39 (2014). 10P. Jedlicka, Front Mol Neurosci. 10, 366 (2017). 11P.D. Bruza, Z. Wang, J.R. Busemeyer J, Trends Cogn. Sci. 19, 383 (2015). 12J.R. Busemeyer, P. Fakhari, P. Kvam, Prog. Biophys. Mol. Biol. 130, 53 (2017). 13A. Louveau et al., Nature Neuroscience 21, 1380 (2018). 14A. Louveau et al., Nature 523, 337 (2015). 15C.W. Lynn, D.S. Bassett, Nature Reviews Physics 1, 318 (2019). 33 16K. Batista-Garca-Ram1 and C.I. Fern´andez-Verdecia, Behav Sci 8(4), 39 (2018). 17B. V´azquez-Rodr´ıguez et al., Proceedings of the National Academy of Sciences, 201903403 (2019). 18A. Mess´e, D. Rudrauf, H. Benali and G. Marrelec, PLoS Comput Biol 10(3), e1003530 (2014). 19F.A. Azevedo, L.R. Carvalho, L.T. Grinberg, J.M. Farfel, R.E. Ferretti, R.E. Leite, F.W. Jacob, R. Lent, S. Herculano-Houzel, J Comp Neurol. 513(5), 532 (2009). 20A. Araque and M. Navarrete, Philos Trans R Soc Lond B Biol Sci. 365(1551), 2375 (2010). 21A. Tameem, H. Krovvidi, Continuing Education in Anaesthesia Critical Care & Pain 13(4), 113 (2013). 22A.L. Hodgkin and A.F. Huxley, J Physiol 117(4), (1952). 23A.E. Pereda, Nat Rev Neurosci. 15(4), 250 (2014). 24H. Lodish, A. Berk, S.L. Zipursky, P. Matsudaira, D. Baltimore and J. Darnell, Molecular Cell Biology, 4th edition, W.H. Freeman New York (2000). 25F. Cavanna, M.G. Vilas, M. Palmucci, E. Tagliazucchi, NeuroImage 180, 383 (2018). 26A.E. Cavanna, A. Nani, H. Blumenfeld, S. Laureys, Neuroimaging of Consciousness, Springer Berlin (2013). 27A.M. Owen, Annual Review of Psychology 64, 109 (2013). 28B. Rohaut, A. Eliseyev and J. Claassen, Critical Care 23, (2019). 29M. Boly, M. Massimini, N. Tsuchiya, B.R. Postle, C. Koch and G. Tononi, Journal of Neuroscience 37(40), 9603 (2017). 30J. Aru, M. Suzuki, R. Rutiku, M.E. Larkum and T. Bachmann, Front. Syst. Neurosci. 13, 43 (2019). 31D. Godwin, R.L. Barry and R. Marois, PNAS 112(12), 3799 (2015). 32G. Tononi, Biol Bull. 215(3), 216 (2008). 33G. Tononi, M. Boly, M. Massimini, C. Koch, Nat. Rev. Neurosci. 17, 450 (2016). 34X.D. Arsiwalla1 and P. Verschure, Front Neurosci. 12, 424 (2018). 35A. de Sousa, Mens Sana Monogr. 11(1), 100 (2013). 36S.R. Hameroff and R. Penrose, Math Comput Simul 40 453 (1996). 37S.R. Hameroff and R. Penrose, J Conscious Stud 3(1), 36 (1996). 38S. Meunier, I. Vernos, Journal of Cell Science 125, 2805 (2012). 34 39M.J. Shelley, Annual Review of Fluid Mechanics 48, 487 (2016). 40P. Dustin, Microtubules (2nd edition), Springer-Verlag, New York (1985). 41G. Guillaud, C. Bosc, A. Fourest-Lieuvin, E. Denarier, F. Priollet, L. Lafanach`ere, D. Job, Cell Biol 142, 167 (1998). 42T. Craddock, J. Tuszynski, S. Hameroff, PLoS Comput Biol 8(3), e1002421 (2012). 43T.J. Craddock, D. Friesen, J. Mane, S. Hameroff, J. Tuszynski, J R Soc Interface 6;11(100), 20140677 (2014). 44J.T. Toole, P. Kurian, T.J.A. Craddock, Journal of Cognitive Science 19, 115 (2018). 45S. Hameroff, J Biol Phys. 36(1), 71 (2010). 46S. Hameroff, Cogn Sci 31, 1035 (2007). 47M. Tegmark, Physical review. E 61(4), 4194 (2000). 48S. Hagan, S. Hameroff, J. Tuszynski, Phys Rev E 65, 061901 (2001). 49G.S. Engel et al., Nature 446, 782 (2007). 50T. Brixner, J. Stenger, H.M. Vaswani, M. Cho, R.E. Blankenship and G.R. Fleming, Nature 434, 625 (2005). 51R. van Grondelle and V.I. Novoderezhkin, Proc. Chem. 3, 198 (2011). 52G.S. Schlau-Cohen et al., J. Phys. Chem. B 113, 15352 (2009). 53G. Panitchayangkoon et al. Proc. Natl. Acad. Sci. USA 107, 12766 (2010). 54E. Collini et al. Nature 463, 644 (2010). 55E. Mandelkow, Y.H. Song, E.M. Mandelkow, Trends Cell Biol 5(7), 262 (1992). 56L.K. McKemmish, J.R. Reimers, R.H. McKenzie, A.E. Mark, N.S. Hush, Phys Rev E 80, 021912 (2009). 57J.R. Reimers, L.K. McKemmish, R.H. McKenzie, A.E. Mark, N.S. Hush, Proc Natl Acad Sci USA 106(11), 4219 (2009). 58J. Pokorn´y, Bioelectrochemistry 63, 321 (2004). 59S. Sahu, S. Ghosh, B. Ghosh, K. Aswani, K. Hirata, D. Fujita, A. Bandyopadhyay, Biosens Bioelectron 47, 141 (2013). 60S. Sahu, S. Ghosh, K. Hirata, D. Fujita, A. Bandyopadhyay, Phys Lett 102, 123701 (2013). 61D. Emerson, B. Weiser, J. Psonis, Z. Liao, O. Taratula, A. Fiamengo, X. Wang, K. Sugasawa, A.B. Smith, R.G. Eckenhoff and I.J. Dmochowski, J Am Chem Soc 135(14), 5398 (2013). 35 62M. Bianchi, A.J. Shah, K.C. Fone, A.R. Atkins, L.A. Dawson, C.A. Heidbreder, M.E. Hows, J.J. Hagan, C.A. Marsden, Synapse 63, 359 (2009). 63A. Rinaldi, EMBO Rep. 15(11), 1113 (2014). 64T.J.A. Craddock, S.R. Hameroff, A.T. Ayoub, M. Klobukowski and J.A. Tuszynski, Cur- rent Topics in Medicinal Chemistry 15, 523 (2015). 65L. Turin, E.M.C. Skoulakis and A.P. Horsfield, PNAS 111(34), E3524 (2014). 66S.R. Hameroff, R.C. Watt, J.D. Borel and G. Carlson, Physiol. Chem. Phys. 14, 183 (1982). 67S.R. Hameroff and R.C. Watt, Anesth. Analg. 62, 936 (1983). 68S.R. Hameroff, Toxicol. Lett. 100-101, 31 (1998). 69S.R. Hameroff, Anesthesiology 105(2), 400 (2006). 70N. Li et al. Anesthesiology 129(2), 271 (2018). 71L.J. Voss, P.S. Garcia, H. Hentschke and M.I. Banks, Anesthesiology 130(6), 1049 (2019). 72C.D. Richards, BJA: British Journal of Anaesthesia 89(1), 79 (2002). 73M. Perouansky, Anesthesiology 117, 465 (2012). 74E.F. Keller, L.A. Segel, J. Theor. Biol. 26, 399 (1970). 75T. Craddock et al., PloS one 7, e37251 (2012). 76R.G. Eckenhoff, J.S. Johansson, Pharmacol Rev. 49(4), 343 (1997). 77R.G. Eckenhoff, Anesth Analg. 107(3), 859 (2008). 78J.M. Sonner, R.S. Cantor Annu Rev Biophys 42, 143 (2013). 79J.Z. Pan, J. Xi, M.F. Eckenhoff, R.G. Eckenhoff, Proteomics 8, 2983 (2008). 80T. Craddock, A. Priel and J. Tuszynski, Journal of integrative neuroscience 13, 293 (2014). 81G. Celardo, M. Angeli, P. Kurian and T. Craddock, New Journal of Physics 21, (2018). 82T.J.A. Craddock, P. Kurian, J. Preto, K. Sahu, S.R. Hameroff, M. Klobukowski and J.A. Tuszynski, Scientific Reports 7(9877), (2017). 83A. Albert, Selective Toxicity: The Physico-Chemical Basis of Therapy, Chapman & Hall, London (1985). 84M.P.A. Fisher, Annals of Physics 362, (2015). 85M.W. Swift, C.G. Van de Walle and M.P.A. Fisher, Physical Chemistry Chemical Physics 20(18), (2017). 36 86C.P. Weingarten, P.M. Doraiswamy and M.P.A. Fisher, Frontiers in Human Neuroscience 10, (2016) 87D.M. Eigler, E.K. Schweizer, Nature 344, 524 (1990). 88C.J. Rourk, Biosystems 171, 48 (2018). 89L. Turin and E.M.C. Skoulakis, Methods Enzymol. 603, 115 (2018). 90J.M. Kater, Science 82(2124), 256 (1935). 91P.J. Cohen, Anesthesiology 39(2), 153 (1973). 92S.A. Pshenichnyuk and A. Modelli, Phys Chem Phys 15(23), 9125 (2013). 93A. Gaitanidis, A. Sotgiu, L. Turin, arXiv:1907.04764 [physics.bio-ph] (2019). 94R.K. Burdick, J.P. Villabona-Monsalve, G.A. Mashour and T. Goodson III, Scientific Reports 9(11351), (2019). 95J. Pan et al., Translational Psychiatry 8(130), (2018). 96D.J. Nutt, J Clin Psychiatry 69,4 (2008). 97A. Vaziri and M.B. Plenio, New Journal of Physics 12, (2010) 98E. Fischer, Ber. Dtsch. Chem. Ges.27, 2985 (1894). 99A. Tripathi and V.A. Bankaitis, J Mol Med Clin Appl. 2(1), (2017). 100H. Mitsuda et al., Journal of Physical Chemistry Letters 1(7), 1130 (2010). 101A. Reese, N. Holmgaard List, J. Kongsted, I.A. Solovyov, PLOS ONE 11(3), e0152345 (2016). 102M.D.H. Samee, B.G. Bruneau, K.S. Pollard, Cell Systems 8(1), 27 (2019). 103M. Rask-Andersen, M.S. Alm´en, H.B. Schioth, Nat. Rev. Drug Discov. 10, 579 (2011). 104A. Gehrckens, A. Horsfield, E. Skoulakis and L. Turin, preprint on biorxiv 10.1101/650531, (2019). 105A.P. Horsfield, A. Haase and L. Turin, Advances in Physics: X 2:3, 937 (2017). 106G.M. Dyson, J. Soc. Chem. Industry 57, 647 (1938). 107L. Turin, Chemical Senses 21(6), 773 (1996). 108Franco MI1, Turin L, Mershin A, Skoulakis EM, Proc Natl Acad Sci U S A 108(9), 3797 (2011). 109R.D. Hoehn, D. Nichols, H. Neven and S Kais, Sci Rep. 5, 9990 (2015). 110J. Hara, Experientia 33(5), 618 (1977). 111B.R. Havens and C.D. Melone, Dev. Food. Sci. 37, 497 (1995) 112K.D. Klika, ISRN Org Chem, 515810 (2013). 37 113M. Paoli, D. Mnch, A. Haase, E. Skoulakis, L. Turin and C. G. Galizia, eNeuro 4(3), 0070 (2017). 114H.K Chee and S.J. Oh, Genomics Inform 11(4), 282 (2013). 115H.K. Chee, J. Yang, J. Joung, B. Zhang, S.J. Oh, FEBS Letters 589(4), 548 (2015). 116Ross D. Hoehn, David E. Nichols, John D. McCorvy, Hartmut Neven, and Sabre Kais, PNAS 114(22), 5595 (2017). 117S. Lin, L. Lee and Y.K. Yang, Clin Psychopharmacol Neurosci. 12(3), 196 (2014). 118M. Krzan, R. Vianello, A. Marsavelski, M. Repic, M. Zaksek, K. Kotnik, E. Fijan and J. Mavri, PLoS One 11(5), e0154002 (2016). 119S.J. Oh, Genomics Inf. 10(2), 128 (2012). 120K. Palczewski, Annu Rev Biochem. 75, 743 (2007). 121A. Krishnan, M.S. Alm´en, R. Fredriksson, H.B. Schioth, PLoS One, 7(1), e29817 (2012). 122D. Wacker, R.C. Stevens and B.L. Roth, Cell 170(3), 414 (2017). 123D. Fels, PLoS One 4, e5086 (2009). 124M. Rahnama, J.A. Tuszynski, I. B´okkon, M. Cifra, P. Sardar, J. Salari, Integr Neurosci. 10(1), 65 (2011). 125S. Kumar, K. Boone, J. Tuszy´nski, P. Barclay and C. Simon, Scientific Reports 6, 36508 (2016). 126M. Cifra, P.J. Posp´ısil, Photochem Photobiol B 139, 2 (2014). 127M. Kobayashi, M. Takeda, T. Sato, Y. Yamazaki, K. Kaneko, K. Ito, H. Kato, Humio Inaba, Neurosci. Res. 34, 103 (1999). 128Y. Isojima, T. Isoshima, K. Nagai, K. Kikuchi and H. Nakagawa, NeuroReport 6, 658 (1995). 129Y. Kataoka et al., Biochem. Biophys. Res. Commun. 285, 1007 (2001). 130R. Tang and J. Dai, PLoS One 9, e85643 (2014). 131Z. Wang, N. Wang, Z. Li, F. Xiao and J. Dai, PNAS 113, 8753 (2016). 132V. Salari, I. B´okkon, R. Ghobadi, F. Scholkmann and J.A. Tuszynski, PNAS 113, E5540 (2016). 133Y. Sun, C. Wang, J. Dai, Photochem Photobiol Sci. 9(3), 315 (2010). 134P. Zarkeshian, S. Kumar, J. Tuszynski, P. Barclay, C. Simon, Front Biosci (Landmark Ed). 23, 1407 (2018). 135A.I. Zhuravlev, O.P. Tsvylev and S.M. Zubkova, Biofizika 18, 1037 (1973). 38 136J.A. Tuszy´nski and J.M. Dixon, Phys. Rev. E 64, 051915 (2001). 137G. Albrecht-Buehler, Cell Motil. Cytoskeleton 27, 262 (1994). 138M. Kato, K. Shinzawa and S. Yoshikawa, Photobiochem. Photobiophys. 2, 263 (1981). 139P. Kurian, T.O. Obisesan, and T. Craddock, Journal of Photochemistry and Photobiology B: Biology 175, (2017). 140L. Tonello, M. Cocchi, F. Gabrielli, J.A. Tuszynski, J Integr Neurosci. 14(3), 295 (2015). 141L. Tonello, B. Gashi, A. Scuotto, G. Cappello, M. Cocchi, F. Gabrielli and J.A. Tuszynski, Journal of Integrative Neuroscience 17, 177 (2019). 142S. Baghbanzadeh and I. Kassal, J. Phys. Chem. Lett. 7, 3804 (2016). 143M.W. Gray, G. Burger and B.F. Lang, Science 283, 1476 (1999). 144Z. Wang and M. Wu, Sci. Rep. 5, 7949 (2015). 145A. Harish and C.G. Kurland, Journal of Theoretical Biology 434, 88 (2017). 146J. Martijn, J. Vosseberg, L. Guy, P. Offre and T.J.G. Ettema, Nature 557, 101 (2018). 147P. Husen, I.A. Solov'yov, Journal of Physical Chemistry B 121 3308 (2017). 148A.B. Salo, P. Husen, I.A. Solov'yov, Journal of Physical Chemistry B 121, 1771 (2017). 149A.M. Barragan, K. Schulten, I.A. Solov'yov, Journal of Physical Chemistry B 120, 11369 (2016). 150D. Xia, L. Esser, W. Tang, F. Zhou, Y. Zhou, L.Yu, C. Yu, Biochimica et Biophysica Acta (BBA) - Bioenergetics 1827, 1278 (20130. 151F. Salehpour et al., Brain Research Bulletin 144, 213 (2019). 152M.R. Hamblin, BBA Clinical 6, 113 (2016). 153M.R. Hamblin, J Neurosci Res. 96(4), 731 (2018). 154J. Chang, Y. Ren, R. Wang, C. Li, Y. Wang, X. Chu, Xiang-Ping. Neuropsychiatry 08, (2018). 155D.W. Barrett, F. Gonzalez-Lima, Neuroscience 230, 13 (2013). 156N.J. Blanco, W.T. Maddox and F. Gonzalez-Lima, J Neuropsychol. 11(1), 14 (2017). 157N.J.Blanco, C.L.Saucedo, F.Gonzalez-Lima, Neurobiology of Learning and Memory 139, 69 (2017). 158T.I. Karu and S.F. Kolyakov, Photomed Laser Surg. 23(4), 355 (2005). 159K. Schulten, C.E. Swenberg and A. Weller Z. Phys. Chem. 111. 1 (1978). 160W. Wiltschko and R. Wiltschko, J Comp Physiol A 191, 675 (2005). 161C.T. Rodgers and P.J. Hore, PNAS 106(2), 353 (2009). 39 162T. Ritz, T. Yoshii, C. Helfrich-Forster and M. Ahmad, Communicative and Integrative Biology 3, 24 (2010). 163A. Pinzon-Rodriguez, S. Bensch, R. Muheim, J. R. Soc. Interface 15, 20180058 (2018). 164A. Gunther, A. Einwich, E. Sjulstok, R. Feederle, P. Bolte, K.W. Koch, I.A. Solov'yov, H. Mouritsen, Curr Biol. 28, 211 (2018). 165M. Liedvogel and H. Mouritsen, J. R. Soc. Interface 7, S147 (2014). 166C.X. Wang, I.A. Hilburn, D. Wu, Y. Mizuhara, C.P. Coust´e, J.N.H. Abrahams, S.E. Bernstein, A. Matani, S. Shimojo and J.L. Kirschvink, eNeuro 6(2), 0483 (2019). 167R. Wiltschko and W. Wiltschko, Commun Integr Biol. 2(2), 100 (2009). 168D.A. Kishkinev, N.S. Chernetsov, Zh Obshch Biol. 75(2), 104 (2014). 169D.S. Hsu et al., Biochemistry 35, 13871 (1996). 170L.E. Foley, R.J. Gegear and S.M. Reppert, Nat Commun. 2, 356 (2011). 171R. Van Wijk, S. Bosman, J. Ackerman, E.P.A. Van Wijk, NeuroQuantology 6, 452 (2008). 172C. Schweitzer and R. Schmidt, Chem. Rev. 103, 1685 (2003). 173A. Marais, I. Sinayskiy, F. Petruccione and R. van Grondelle, Scientific Reports 5, 8720 (2015). 174R.J. Usselman, C. Chavarriaga, P.R. Castello, M. Procopio, T. Ritz, E.A. Dratz, D.J. Singel and C.F. Martino, Scientific Reports 6, 38543 (2016). 175R.J. Usselman , I. Hill, D.J. Singel, C.F. Martino, PLoS ONE 9(3), e93065 (2014). 176H. Wang and X. Zhang, Int J Mol Sci. 18(10), 2175 (2017). 177M.K. Shigenaga, T.M. Hagen,and B.N. Ames, Proc. Natl. Acad. Sci. 91, 10771 (1994). 178E.R. Stadtman and B.S. Berlett, Chem. Res. Toxicol. 10, 485 (1997). 179T.M. Michel, D. Pulschen, J. Thome, Curr Pharm Des. 18(36), 5890 (2012). 180S. Salim, J Pharmacol Exp Ther. 360(1), 201 (2017). 181S. Salim, Curr Neuropharmacol. 12(2), 140 (2014). 182P. Kovacic and W. Weston, Chronicles of Pharmaceutical Science 1(6), (2017). 183P. Thalau, T. Ritz, K. Stapput, R. Wiltschko and W. Wiltschko, Naturwissenschaften 92, 86 (2004). 184S. Engels, N.L. Schneider, N. Lefeldt, C.M. Hein, M. Zapka, A. Michalik, D. Elbers, A. Kittel, P.J. Hore and H. Mouritsen, Nature 509, 353 (2014). 185J. Juutilainen, M. Herrala, J. Luukkonen, J. Naarala, P.J. Hore, Proc Biol Sci. 285(1879), 20180590 (2018). 40 186P.J. Hore, eLife 8, (2019). 187L.M. Lyell et al., Proc Biol Sci. 285(1879), 20180590 (2018). 188P. Boyce, E.A. Barriball, Fam Physician 39(5),307 (2010). 189A. Ferguson et al., EBioMedicine 35, 279 (2018). 190J. Close, Proc Biol Sci. 279(1736), 2081 (2012). 191R.W. Kay, Br. J. Psychiatry 164, 403 (1994). 192V.N. Oraevskii, V.P. Kuleshova, I.F. Gurfinkel, A.V. Guseva, S.I. Rapoport, Biofizika 43, 844 (1998). 193M.B. Plenio and S.F. Huelga, New J. Phys. 10, 113019 (2008). 194G.S. Engel, Procedia Chemistry 3, 222 (2011). 195H. Hu and M. Wu, Medical Hypotheses 63, 633 (2004). 196L.A. Cole, Biology of Life: Biochemistry, Physiology and Philosophy Chapter 10, 65 (2016). 197A.S. Posner and F. Betts, Acc Chem Res. 8(8), 273 (1975). 198T.C Player and P.J. Hore, J. R. Soc. Interface 15(147), (2018). 199L. Wang, S. Li, E. Ruiz-Agudo, C.V. Putnis, A. Putnis, Cryst. Eng Comm. 14(19), 6252 (2012). 200G. Mancardi, C.E.H. Tamargo, T.D. Di, N.H. de Leeuw, J Mater Chem B. 5(35), 7274 (2017). 201G. Stevanato et al., Angew Chem Int Ed. 54(12), 3740 (2015). 202K. Onuma and A. Ito, Chem. Mater. 10, 3346 (1998). 203J.A. Sechzer et al., Biol. Psychiatry 21, 1258 (1986). 204N.Y Halpern and E. Crosson, Annals of Physics 407, 92 (2019). 205C. Marletto et al., J. Phys. Commun. 2, 101001 (2018). 206B.E. Kane, Nature 393, 133 (1998). 207G. Tosi, F.A. Mohiyaddin, V. Schmitt, S. Tenberg, R. Rahman, G. Klimeck and A. Morello, Nature Communications 8, (2017). 208Y. He, S.K. Gorman, D. Keith, L. Kranz, J. G. Keizer and M.Y. Simmons, Nature 571, 371 (2019). 209C. Kloeffel and D. Loss, Annual Review of Condensed Matter Physics 4, 51 (2013). 210C.W. Lee, H. Zhang, L. Geng and H.B. Peng, PLoS One 9(2), e90187 (2014). 211D. Kim et al., Nature Nanotechnology 13, 812 (2018). 41 212Y. Liu, L. Xu, W. Dai, H. Dong, Y. Wen and X. Zhang, Nanoscale,7, 19060 (2017). 213V.M. Mazhul, and D.G. Shcherbin, Biofizika 44, 676 (1999). 42
1209.3829
1
1209
2012-09-18T02:25:50
Self-organized criticality in a network of interacting neurons
[ "q-bio.NC", "physics.bio-ph" ]
This paper contains an analysis of a simple neural network that exhibits self-organized criticality. Such criticality follows from the combination of a simple neural network with an excitatory feedback loop that generates bistability, in combination with an anti-Hebbian synapse in its input pathway. Using the methods of statistical field theory, we show how one can formulate the stochastic dynamics of such a network as the action of a path integral, which we then investigate using renormalization group methods. The results indicate that the network exhibits hysteresis in switching back and forward between its two stable states, each of which loses its stability at a saddle-node bifurcation. The renormalization group analysis shows that the fluctuations in the neighborhood of such bifurcations have the signature of directed percolation. Thus the network states undergo the neural analog of a phase transition in the universality class of directed percolation. The network replicates precisely the behavior of the original sand-pile model of Bak, Tang & Wiesenfeld.
q-bio.NC
q-bio
Self-organized criticality in a network of interacting neurons J D Cowan1, J Neuman2, W van Drongelen3 1 Dept. of Mathematics, University of Chicago, 5734 S. University Ave., Chicago, IL 60637 2 Dept. of Physics, University of Chicago, 5720 S. Ellis Ave., Chicago, IL, 60637 3 Dept. of Pediatrics, University of Chicago, KCBD 900 E. 57th St., Chicago, IL., 60637 E-mail: [email protected] Abstract. This paper contains an analysis of a simple neural network that exhibits self-organized criticality. Such criticality follows from the combination of a simple neural network with an excitatory feedback loop that generates bistability, in combination with an anti-Hebbian synapse in its input pathway. Using the methods of statistical field theory, we show how one can formulate the stochastic dynamics of such a network as the action of a path integral, which we then investigate using renormalization group methods. The results indicate that the network exhibits hysteresis in switching back and forward between its two stable states, each of which loses its stability at a saddle-node bifurcation. The renormalization group analysis shows that the fluctuations in the neighborhood of such bifurcations have the signature of directed percolation. Thus the network states undergo the neural analog of a phase transition in the universality class of directed percolation. The network replicates precisely the behavior of the original sand-pile model of Bak, Tang & Wiesenfeld. 2 1 0 2 p e S 8 1 ] . C N o i b - q [ 1 v 9 2 8 3 . 9 0 2 1 : v i X r a Self-organized criticality in a neural network 2 1. Introduction The idea of self-organized criticality (SOC) was introduced by Bak et al. (1988). Their paper immediately triggered an avalanche of papers on the topic, not the least of which was a connection with 1/f - or scale-free noise. However it was not until another paper appeared, by Gil & Sornette (1996), which greatly clarified the dynamical prerequisites for achieving SOC, that a real understanding developed of the essential requirements for SOC: (1) an order-parameter equation for a dynamical system with a time-constant τo, with stable states separated by a threshold, (2) a control-parameter equation with a time-constant τc, and (3) a steady driving force. In Bak et.al.'s classic example, the sand-pile model, the order parameter is the rate of flow of sand grains down a sand-pile, the control parameter is the sand-pile's slope, and the driving force is a steady flow of grains of sand onto the top of the pile. Gil and Sornette showed that if τo (cid:28) τc then the resulting avalanches of sand down the pile would have a scale-free distribution, whereas if τo (cid:29) τc then the distribution would also exhibit one or more large avalanches. In this paper we will analyze a neural network model which is in one-to-one correspondence with the Gil-Sornette SOC-model, and therefore also exhibits SOC. 1.1. Neural network dynamics Consider first the mathematical representation of the dynamics of a neocortical slab comprising a single spatially homogeneous network of N excitatory neurons. Such neurons make transitions from a quiescent state q to an activated state a at the rate σ and back again to the quiescent state q at the rate α, as shown in figure 1. Figure 1. Neural state transitions Let Pn(t) be the probability that a fraction n/N is active at time t. Then the probabilistic dynamics of such a slab can be formulated as a master equation of the Self-organized criticality in a neural network form: dPn(t) dt = α[(n + 1)Pn+1 − nPn] + (N − n + 1)f [s(n − 1)]Pn−1 − (N − n)f [s(n)]Pn 3 (1) where α is the rate at which activated neurons become quiescent, and s(n) is the total current or excitation driving each neuron in the population to fire at the rate σ = f [s(n)]. We assume in this simplified model that each neuron receives a signal weighted by w0/N from each of N other neurons in the population, so that s(n) = w0n + h where h is an external current. (2) Equation 1 can be extended to the spatially inhomogeneous case. Let nr/Nr be the fraction of active cells at time t in the rth population of Nr cells, and let P [n, t] be the probability of the configuration n = {n1, n2,··· , nr,··· , nΩ} existing at time t. The extended master equation then takes the form (cid:88) (cid:88) r dP [n, t] dt = α [(nr + 1)P [nr+, t] − nrP [n, t]] [(Nr − nr + 1)f [s(nr − 1)]P [nr−, t] + − (Nr − nr)f [s(nr)]P [n, t]] r (4) (5) (6) (7) (3) where nr± = {n1, n2,··· , nr ± 1,··· , nΩ} and there are a total of Ω locally homogeneous populations. Following Van Kampen (1981) equation 3 can be rewritten using the one-step operators i.e., E± r f (n) = f (nr ± 1) (cid:88) (cid:2)α(E + dP [n, t] dt = r − 1)nr + (E− r − 1)(Nr − nr)f [s(nr)](cid:3) P [n, t] Note that the total number of cells Nr in the rth population comprises nr active cells and Nr − nr = qr quiescent cells, so that nr + qr = Nr Thus equation 5 can be rewritten slightly in the form (cid:88) (cid:2)α(E + dP [n, t] dt = r − 1)nr + (E− r − 1)qrf [s(nr)](cid:3) P [n, t] r r Self-organized criticality in a neural network 4 1.2. Annihilation and creation operators We now introduce Fock space annihilation and creation operators satisfying boson commutation rules [ar, a† [ar, as] = [a† [qr, qs] = [q† r, a†s] = 0 r, q† s] = 0 s] = δ(r − s) s] = [qr, q† (8) such that given neural vectors nr(cid:105), mr(cid:48)(cid:105) rnr(cid:105) = nr + 1(cid:105), arnr(cid:105) = nrnr − 1(cid:105) a† rmr(cid:105) = mr + 1(cid:105), qrmr(cid:105) = mrmr − 1(cid:105) q† where r0r(cid:105),mr(cid:105) = (q† r 0r(cid:105) r)m and 0r(cid:105) is a fiducial or vacuum (empty) state such that nr + mr = 1,nr(cid:105) = (a†)n r0r(cid:105) = nr = 1r(cid:105), ar0r(cid:105) = 0 a† r0r(cid:105) = mr = 1r(cid:105), qr0r(cid:105) = 0 q† and there exists a dual vector (cid:104)0 such that r = 0, r = 0, (cid:104)0rar = (cid:104)nr = 1r,(cid:104)0ra† (cid:104)0rqr = (cid:104)mr = 1r,(cid:104)0rq† It follows from equation 9 that rarnr(cid:105) = nrnr(cid:105) a† (9) (10) (11) (12) i.e. nr is the eigenvalue of the operator a† or number density operator. rar, which is therefore referred to as a number 1.3. Bosonizing the master equation The master equation can now be bosonized by replacing the van Kampen one-step operators by bosonic equivalents, i.e. r)ar r)qr (E + r − 1) = (q† r − a† r − 1) = (a† (E− r − q† (cid:88) so eqn 7 becomes: (cid:88) r)ar + (a† Let P (t)(cid:105) be a probability state vector satisfying (cid:2)α(q† r − a† dP [n, t] dt = r P (t)(cid:105) = P [n, t]n(cid:105) n rar)](cid:3) P [n, t] r − q† r)qrf [s(a† (13) (14) (15) P (t)(cid:105) = r − a† r)ar + (a† r − q† r)qrf [s(a† rar)](cid:3)P (t)(cid:105) Self-organized criticality in a neural network Then equation 7 can be written in the form (cid:88) (cid:2)α(q† r or formally as d dt d dt where − H = P (t)(cid:105) = − HP (t)(cid:105) (cid:88) (cid:2)α(q† r − a† r)ar + (a† r − q† r)qrf [s(a† rar)](cid:3) 5 (16) (17) (18) (20) (21) is the quasi-Hamiltonian operator for the Markov process represented in equation 1. r 1.4. From bosons to coherent states Equation 18 is a linear operator equation with formal solution P (t)(cid:105) = exp[− H(t − t0)P (t0)(cid:105) We need to re-express this solution in terms of numbers rather than operators. This can be achieved by introducing coherent states. These were introduced by Schrodinger (1926) and first used extensively in coherent optics by Glauber (1963). We therefore introduce such states φr(cid:105) in the form φr(cid:105) = exp[−1 2 rϕr + ϕra† ϕ(cid:63) r]0r(cid:105) (19) where ϕr is the right eigenvalue of ar, i.e. arφr(cid:105) = ϕrφr(cid:105). There is also a coherent state representation of qr in the form θr(cid:105) such that the right eigenvalue of qr is ϑr, i.e. qrθr(cid:105) = ϑrθr(cid:105). In similar fashion (cid:104)φra† r = (cid:104)φr ϕr where ϕr, the complex conjugate of r = (cid:104)θr ϑr, i.e. ϑr is the left eigenvalue ϕ, is the left eigenvalue of a† of q† r, and similarly (cid:104)θrq† All this suggests that the operator quasi-Hamiltonian has a coherent state representation in the form rarφr(cid:105) = (cid:104)φr ϕrϕrφr(cid:105) = ϕrϕr (cid:88) (cid:104) − H = α( ϑr − ϕr)ϕr + ( ϕr − ϑr)ϑrf [s( ϕrϕr)] (cid:105) r. It follows that (cid:104)φra† r Note in passing that operator products involving powers of the number operator a† rar must first be normal ordered, i.e. all creation operators a† r must preceed the annihilation operators ar, before coherent states can be introduced. For example the normal order form of exp[a†a] written as : exp[a†a] : is expanded as : exp[a†a] : = 1 + (a†a) + = 1 + (a†a) + (a†a)2 + ··· (a†a + a†2a2) + ··· 1 2! 1 2! sl,ja†jaj ∞(cid:88) l(cid:88) 1 l! l=0 j=0 = (22) Self-organized criticality in a neural network 6 where the sl,j are Stirling numbers of the second kind. It follows that : exp[a†a] : can be written as : exp[a†a] := hka†kak (23) (cid:88) k=0 where hk =(cid:80) l sl,k. The final preliminary step in this formulation is to take the continuum limit of the expression for H in equation 21, so that (cid:90) (cid:90) (cid:104) − H = (24) where ϕr → ϕ(x, t) ≡ ϕ etc., and the conjugate coherent state ϕ has been shifted to 1 + ϕ. α( ϑ − ϕ)ϕ + ( ϕ − ϑ)ϑf [s( ϕϕ + ϕ)] ddxdt (cid:105) 1.5. Dimensions and the density representation Before proceeding further we need to assign a dimension to each variable in equation 24. To do so we use a modified version of the convention used in particle physics so that [x] = L−1, [t] = L−2 where L is the length scale used, whence [x2/t] = L0. {This generates a scaling found in Markov random walks and related processes such as stochastic neural activity.} Then [α] = L2, [ ϕϕ] = Ld, [f [s]] = [α] = L2. This last value of [f [s]] implies that the input current function s( ϕϕ + ϕ) = s(I) = kI where the constant k has the dimensions of inverse current. The net effect of such a choice leads to the required result that [H] = 0. [ϕ] = Ld, [ ϕ] = L0, To emphasize this choice we further transform the coherent-state quasi-Hamiltonian by introducing the density representation: ϕ → exp[n] − 1, ϕ → n[exp[−[n]] ϑ → exp[p] − 1, ϑ → p[exp[−[p]] so that equation 24 transforms into (cid:90) (cid:90) − H = ddxdt [α(exp(p − n) − 1)n + (exp(n − p) − 1)pf [s(n)]] (26) k(w(cid:63)n+h) where (cid:63) is the spatial convolution operator, i.e. w(cid:63)n =(cid:82) ddx(cid:48)w(x−x(cid:48))n(x(cid:48), t). Note that in the continuum limit the input current function s(n) = s(I) = kI = (25) 1.6. From the quasi-Hamiltonian to a neural Path Integral Using standard methods (Doi 1976, Peliti 1985) Buice & Cowan (2007) incorporated the quasi-Hamiltonian into the action of a Wiener path integral. This action takes the form: (cid:90) (cid:90) S(n) = ddxdt [n∂tn + p∂tp+ α(1 − exp(−(n − p))n − (exp(n − p) − 1)pf [s(n)]](27) Self-organized criticality in a neural network 7 The utility of this action is that it is part of the exponent of the moment generating functional for the statistical moments of the probability density P [n, t]. At an extremum (cid:12)(cid:12)(cid:12)(cid:12)n=0 (cid:12)(cid:12)(cid:12)(cid:12)p=0 δS(n) δn = δS(n) δ p = 0 which leads to the mean-field equations ∂tn + αn − pf [s(n)] = 0, ∂tp − αn + pf [s(n)] = 0 whence n + p = ρ (28) (29) (30) where ρ is the (constant) packing density of excitatory neurons in the population. This is a mean-field result, so we replace p by ρ − ncl in equation 27, so as to generate the mean-field Wilson-Cowan equation (Wilson & Cowan 1973) for a single spatially organized population from the first variation of the action given in equation 27 and (in principle), all higher moments, and (finally) rewrite equation 27 in the form: (cid:90) (cid:90) S(n) = ddxdt [n∂tn+ α(1 − exp(−n))n − (exp(n) − 1)(ρ − ncl)f [s(n)]] (31) 1.7. Renormalizing the path integral We expand n about its mean value (cid:104)n(cid:105) = ncl, which satisfies equation 28 in the form: (32) ∂tncl = −αncl + (ρ − ncl)f [s(ncl)] where s(ncl) = k(w (cid:63) ncl + hcl) (33) Thus n → n+ncl, n → n, since ncl = 0. So s(n) → s(n+ncl) and f [s(n)] → f [s(n+ncl)]. If follows that s(n + ncl) = k(w (cid:63) (n + ncl) + (h + hcl)) = k(w (cid:63) ncl + hcl) + k(w (cid:63) n + h)) = s(ncl) + s(n), and therefore f [s(n)] = f [s(ncl) + s(n)]. We next expand f [s(n)] in a Taylor expansion about the mean-field value ncl, noting that from equation A.1, s(n) = k(w (cid:63) n + h)) = k(Ln + h). In what immediately follows we assume that the external stimulus h(x, t) = 0. It follows that: f [s(n)] = f [kL(ncl + n)] = f [kLncl] + f (1)[kLncl]kLn However because of normal ordering, equation 34 leads to the expression: (cid:88) f [s(n)] = gm(kLn)m, where gm = m l=0 f (2)[kLncl](kLn)2 + ··· + 1 2 m(cid:88) f (l) l! sl,m (34) (35) Since the leading terms of gm are proportional to f (m), and given the assumed form for f [s(n)] to be such that f (1) > 0 and f (2) < 0, then gm > 0 for m odd, and gm < 0 for m even. Self-organized criticality in a neural network We also expand the functions exp(±n). The resulting action S(n) takes the form: 8 (cid:90) (cid:90) S(n) = ddxdt [n(∂t + α − (ρ − ncl)g1kL)n (α + (ρ − ncl)g1kL)n + n((ρ − ncl)g2(kL)2)n2 ((ρ − ncl)g2(kL)2)n2 + ··· (cid:21) (36) − n2 2 n2 2 + It follows from the appendix that w2 is small compared to w0, so that in most expressions the terms proportional to ∇2mnm can be neglected. However this is not always the case for m = 1. Thus the first term can be written approximately as 2!w2∇2))n = n(∂t + µ− D∇2)n where µ = α− (ρ− ncl)g1kw0 n(∂t + α− (ρ− ncl)g1k(w0 + 1 2(ρ − ncl)g1kw2. So the expression for the action is now reduced to the form: and D = 1 (cid:90) (cid:90) S(n) = ddxdt(cid:2)n(∂t + µ − D∇2)n − n2G1n (cid:21) +nG2n2 + n2G2n2 + ··· 1 2 (37) where G1 = 1/2(α+(ρ−ncl)g1kw0), and G2 = (ρ−ncl)g2k2w2 that the last term in S(n), i.e., 1 of the renormalization group. 0. We need to demonstrate 2 n2G2n2, and all other terms, are irrelevant in the sense The renormalization group [RG] analysis is carried out via dimensional analysis. It can be shown that all the terms in S(n) are zero-dimensional when [ddxdt] = L−(d+2) and integrated over d-dimensional space and over time, [any term in the integrand] = Ld+2. However, as it stands [n] = Ld, but [n] = L0, so that [nn] = L0+d = Ld. This is not suitable for the scaling analysis implemented in the RG process. We therefore introduce a new scaling, i.e., (cid:114) G1 (cid:114)G2 n, s = s = (38) such that ss = nn where [G2/G1] = L−d. The effect of this scaling is that both s and s have dimension Ld/2. Let G2 n G1 G1G2 = u, G2 = 2τ (39) The net effect of this scaling transformation is that ddxdt(cid:2)s(∂t + µ − D∇2)s + us(s − s)s + τ s2s2 + ···(cid:3) S(s) = (40) But [τ /u] = L−d/2 and therefore scales to zero as L → ∞ under subsequent RG transforms. So asymptotically the terms τ s2s2 + ··· become irrelevant in the RG sense. So finally ddxdt(cid:2)s(∂t + µ − D∇2)s + us(s − s)s(cid:3) S(s) = (41) (cid:90) (cid:90) (cid:90) (cid:90) (cid:112) is the renormalized action of the large-scale neural activity of a single neural population. Self-organized criticality in a neural network 9 1.8. Directed Percolation This action is well-known: it is called Reggeon Field Theory, and is found in directed percolation [DP] in random graphs, in contact processes, in high-energy nuclear physics, in bacterial colonies, all of which exhibit the characteristic properties of what is called a universality class, i.e., it is a phase transition with a universal scaling of important statistical exponents. It also shows up in branching and annihilating random walks, catalytic reactions, and interacting particles. Thus we have mapped the mathematics of large-scale neural activity in a single homogeneous neural population into a percolation problem in random graphs, or equivalently into a branching and annihilating random walk. A first version of this work was presented in Buice & Cowan (2007). A more extensive paper with many applications to neuroscience was presented in Buice & Cowan (2009). Here we note that there is an upper critical dimension at which directed percolation crosses over to mean-field behavior. This upper critical dimension is d = 4. What is the dimension of the neocortex? To answer this question we look at the number of synapses per neuron in the neocortex. Using estimates provided by Stevens (1989), this number is about 4 × 103. Assuming the number of synapses in a terminal axonal arbor to be about 50, the number of neural neighbors per neuron is about 80, so the effective dimensionality of a neocortical hyper lattice is about d = 40. Thus the critical exponents characterizing the neural phase transition are the d = 4 exponents of directed percolation. These have been calculated by Abarbanel & Broznan (1974), Abarbanel et al. (1976), and Amati et al. (1976), and appear in the linear response of the neocortical model to an impulsive stimulus, known to mathematicians as the Green's function and to physicists as the propagator. This takes the form G(x − x(cid:48), t − t(cid:48)) ∝  (t − t(cid:48))−2 exp (cid:105) (cid:104)− x−x(cid:48))2 (cid:105) (cid:104)− x−x(cid:48))2 4(t−t(cid:48)) − µ(t − t(cid:48)) (cid:104)(cid:112)µ(t − t(cid:48)) − x − x(cid:48)(cid:105) 4(t−t(cid:48)) ) (t − t(cid:48))−2 exp µ2Θ , µ > 0 µ = 0 (42) µ < 0 , , where the cases µ > 0, µ = 0, and µ < 0 correspond, respectively, to the sub-critical, critical, and super-critical propagators. They correspond, respectively, to solutions of the cable equation, the diffusion equation, and a nonlinear wave-equation. The critical propagator is thus the diffusion limit of Brownian motion. It turns out that there is a great deal of data supporting the hypothesis that the mean-field propagator of DP correctly describes the essential features of large-scale neocortical activity on many spatio-temporal scales. [See Burns (1951), Lampl et al. (1999), Nauhaus et al. (2009).] In addition data on the statistical structure of large-scale activity recorded in cortical slices by Beggs & Plenz (2003) supports the hypothesis. In particular, the avalanche-size distribution of spontaneous activity in cortical slices fits the DP hypothesis. The analysis can be extended to deal with a neural network comprising both excitatory and inhibitory neurons. However in this paper we describe how to incorporate synaptic plasticity into a network comprising only excitatory neurons, as a mechanism Self-organized criticality in a neural network 10 that tunes the network so that it automatically reaches the critical point of the DP phase transition, thus exhibiting self-organized criticality. 2. Incorporating synaptic plasticity Consider first a single excitatory population model with a fixed recurrent excitatory synapse wE and an input H through an excitatory modifiable synapse wH. The mean- Figure 2. A recurrent excitatory network field neural equation for this is a version of equation 32, i.e. dnE dt = −αEnE + (1 − nE)σE [sE(IE)] where sE(IE) = IE/IRH,E, IE = wE (cid:63) nE + wH (cid:63) nH and the synapse wH is modifiable with dwH dt = −gE(cid:104)(nE − ρE,0 − ρE,SwH) nH(cid:105)t (43) (44) (45) where ρE,0 is a constant neural activity, ρE,S is a constant (Vogels et al. 2011), and gE is a state-dependent rate function. This is also a mean-field equation in which the synaptic weight wH is depressed by an anti-Hebbian mechanism, and potentiated by the input activity nH. The problem is to write an action that incorporates these equations. Note that the time scale of the growth and decay of neural activity is set by the constant αE, whereas that of the growth and decay of synaptic plasticity is set by gE. Thus the ratio αE/gE is an important parameter. 3. Deriving an action for synaptic plasticity The first problem is to develop an action for the modifiable synapse wH. In order to do so we first note from equation A.2 that wH scales with bH, the synaptic conductance or weight, and we can write an approximation to eqn. 45 in the form: dbH dt = −gE (nE − ρE,0 − ρE,SkHbH) nH (46) Self-organized criticality in a neural network 11 We next reformulate the changes in bH as a Markov process with discrete states in continuous time. We therefore assume that bH is quantized in units ∆ of synaptic weight, and similarly for bE. Thus bE = mE∆, bH = mH∆ (47) and we look at the Markov process represented in figure 3: Figure 3. Synaptic state transitions and we write a master equation for this process in the form dP (mH, t) dt = t−(mH + 1)P (mH + 1, t) − t−(mH)P (mH, t) + t+(mH − 1)P (mH − 1, t) − t+(mH)P (mH, t) where P (m, t) is the probability that the synaptic weight mH = m at time t. and t+ = gE(ρE,0 + ρE,SkHmH)nH t− = gEnEnH are the transition rates for the excitatory synapse mH. 3.1. The van Kampen ladder operators We again introduce the van Kampen ladder operators so that the master equation can be rewritten as: − 1) + t+(E− dP (mH, t) mH − 1)(cid:3) P (mH, t) mm = m ± 1 E± =(cid:2)t−(E+ mH dt (48) (49) (50) (51) An examination of eqns. 48-50 indicates that only the transition rate t+ contains a term proportional to mH. We will utilize this property in what follows. Self-organized criticality in a neural network 12 3.2. From the Master Equation to the Action We now introduce bosonic annihilation and creation operators for mH. Let such operators be denoted by s† and s respectively, and let m(cid:105) be a column vector representing the synaptic weight m such that s†m(cid:105) = m + 1(cid:105), sm(cid:105) = mm − 1(cid:105) (52) Note that s† and s satisfy the same commutation rules and equations that we introduced earlier, so that m − 1 → s† − s†s = s†(1 − s) E+ m − 1 → s − s†s = (1 − s†)s E− So we consider the equation: ∂ m(cid:105) ∂t = (E− m − 1)m(cid:105) = (1 − s†)sm(cid:105) We further note that (1 − s†)s is normal ordered. We therefore shift s† and s, s† → 1 + s, s → s where s and s are now coherent states, and introduce the density representation s → exp( m) − 1, s → m exp(− m) we find ∂tm(cid:105) = − ssm(cid:105) = − (exp( m) − 1)m exp(− m)m(cid:105) = − (1 − exp(− m))mm(cid:105) Eqn. 21 tells us that the action Sm must contain a term of the form −gEρE,SkHnH(1 − exp(− mH))mH, a source term −gEρE,0nH mH, and an interaction term of the form (53) (54) (55) (56) (57) (59) (60) leading to an action of the form: +gEnEnH mH, S(mH) = ddxdt [ mH∂tmH − gEρE,SkHnH(1 − exp(− mH))mH − mHgEρE,0nH + mHgEnEnH] (58) Using variational techniques, we can derive the mean-field equation [eqn 45] from the condition: (cid:90) (cid:90) (cid:12)(cid:12)(cid:12)(cid:12) mH =0 δS( mH) δ mH = 0 For then we obtain the equation: dmH dt = −gE (nE − ρE,0 − ρE,SkHmH) nH i.e., eqn. 46. Self-organized criticality in a neural network 13 3.3. Renormalizing the synaptic plasticity action We now proceed to renormalize the action S(mH) just as we renormalized S(nE). We therefore expand the exponential term in equation 58 and rewrite S(mH) in the form: (cid:90) (cid:90) (cid:20) S(mH) = ddxdt mH∂tmH − H1 mHmHnH + −H2 mHnH + mHgEnEnH] 1 2 H1 m2 HmHnH (61) where H1 = gEρE,SkH, and H2 = gEρE,0. We now introduce the scaling (cid:114)H2 H1 (cid:114) H1 H2 mH, sH = sH = (62) such that sHsH − mHmH, and [H1/H2] = L−d. This scaling is analogous to the scaling of n and n which we carried out earlier for neural activities. As before the effect of this scaling is that both sH and sH have dimension Ld/2. Let mH (cid:112) and recall that equation 38 scales nE to(cid:112)G1/G2sE. H1H2 = uH, H1 = 2τH (63) Following the procedure outlined earlier we can calculate which terms in the transformed action S(sH) become irrelevant under scaling transformations. The resulting renormalized synaptic plasticity action takes the form: S(sH) = ddxdt [sH∂tsH − uH sHnH] (64) (cid:90) (cid:90) 4. Combining the actions (cid:90) (cid:90) It follows from this formulation that the full action for the coupled system of equations for the evolution of nE and mH can be obtained simply by adding the actions S(nE) and S(mH) together. The combined action therefore takes the form: ddxdt [nE∂tnE + α(1 − exp(−nE))nE −(exp(nE) − 1)(ρ − nE,cl)f [s(nE)] + mH∂tmH −gEρE,SkHnH(1 − exp(− mH))mH − mHgEρE,0nH + mHgEnEnH] S(nE, mH) = (65) 4.1. A simulation of the behavior of the combined mean-field equations The first variation of equation 65 generates the mean-field equations for nE and mHin the form: dnE dt dmH dt = − αEnE + (1 − nE)σE [sE(IE)] = − gE (nE − ρE,0 − ρE,SkHmH) nH (66) Self-organized criticality in a neural network where sE(IE) = kE(mE (cid:63) nE + mH (cid:63) nH) 14 (67) These equations can be simulated. The results are shown in figure 4. It will be seen that Figure 4. Neural state transitions between a ground state and an excited state. Parameter values: mE = 3, nH = 3; α = 0.2. N (cid:63) is the fixed-point value of nE, and WH is the magnitude of the anti-Hebbian synapse in the input path. in the "ground-state" of low values of N (cid:63) = n(cid:63) E the synaptic weight wH (proportional to mH), increases until it reaches the critical point at a saddle-node bifurcation, at which point N (cid:63) becomes unstable and the system switches to the "excited-state". But then the anti-Hebbian term in the synaptic plasticity dynamics kicks in, and WH declines until the excited-state fixed-point becomes unstable at the upper critical point, also at a saddle-node, and switches back to the ground-state fixed point, following which the hysteresis cycle starts over. This is a exact representation of the sand-pile model's behavior. This is another representation of SOC in a neural network. The reader should compare this with the synaptic mechanisms for achieving SOC described in Levina et al. (2007) and in Millman et al. (2010). 4.2. Renormalizing the combined action To renormalize the combined action we follow the same procedure as before and expand the exponential functions exp[±nE], exp[± mH], and f [s(nE)]. Note that after normal ordering and collecting terms, f [s(nE)] can be expanded in the extended form f [s(nE)] =(cid:80) renormalized combined action takes the form: (cid:90) (cid:90) Self-organized criticality in a neural network 15 m gm(kE(LEnE + LHnH))m. After scaling and dimensional analysis, the ddxdt(cid:2)sE(∂t + µE − DE∇2)sE + uE sE(sE − sE)sE S(sE, sH) = −vE sEnH + sH∂tsH − uH sHnH] (68) where uE, uH and vE are renormalized constants. It would appear that apart from the source term vE sEnH the coupled action is just the sum of the two uncoupled renormalized actions. This is indeed the case! All the addition terms which appear in the current function become irrelevant under renormalization, and do not effect the renormalized action representing nE. Similarly for mH. It follows that fluctuations in the activity nE in the neighborhood of the critical point µE = 0, i.e. those in the fluctuation-driven regime, should be essentially those of directed percolation. 5. Inhibitory synapses In case wH is an inhibitory synapse the transitions t+(m) and t−(m) are reversed. so that: t+ = gEnEnH t− = gE(ρE,0 + ρE,SkHmH)nH (69) Thus now only the transition rate t− contains a term proportional to mH. The effect of this is that the action for Sm leads to the mean-field equation: dmH dt = gE (nE − ρE,0 − ρE,SkHmH) nH at an inhibitory synapse mH. (70) Thus inhibitory feedforward synapses are Hebbian with stimulus dependent depression, whereas excitatory feedforward synapses are anti-Hebbian with stimulus dependent potentiation, and we have now formulated actions for feedforward excitatory and inhibitory synaptic plasticity, based on simple microscopic potentiation and depression processes, involving a unary variable, the synaptic weight mH. 6. Conclusion In this paper we have indicated how one can formulate and analyze actions for a simple network of excitatory cells with an input coupled to the network via an activity- dependent modifiable synapse. This action allows, in principal, the computation of statistical moments of the fluctuating dynamics of the network. Previous work by Wilson & Cowan (1972) indicates that the mean-field dynamics of the network is bistable, and can generate a hysteresis loop. On coupling this dynamics to that of the modifiable feedforward synapse introduced, all the conditions for the achievement of SOC are present in the combined system. The simulation of the mean-field dynamics shows that SOC is indeed achieved. It only remains to simulate the behavior of the combined Self-organized criticality in a neural network 16 system with intrinsic noise. Our prediction is that the fluctuations in the activity near the critical points of the system will exhibit the properties of directed percolation. This will be the subject of another paper. Acknowledgments The work reported in this paper was initially developed in large part with Michael. A. Buice. The current work was supported (in part) by a grant to Wim van Drongelen from the Dr. Ralph & Marian Falk Medical Trust. Appendix Appendix A.1. Expanding the weighting function We approximate the convolution w (cid:63) n as: w (cid:63) n = ddx(cid:48)w(x − x(cid:48))n(x(cid:48), t) (cid:90) w2∇2 + ···)n(x(cid:48), t) (cid:39) (w0 + ≡ Ln 1 2! where w(x) → w(r) = b/σde−r/σ, σ = r0. It follows that w0 = w2 = (cid:90) (cid:90) ddxw(x) = b dΓ(d) Γ(d/2 + 1) πd/2, ddx2w(x) = bσ2 dΓ(d + 2) Γ(d/2 + 1) πd/2 (A.1) (A.2) (A.3) In case d = 3, w0 = 8πb, w2 = 12w0σ2, whence w2/w0 = 12σ2. This expansion of the weighting function w(x) is known as the moment expansion. References Abarbanel H & Broznan J 1974 Phys. Lett. B 48, 345–348. Abarbanel H, Broznan J, Schwimmer A & Sugar R 1976 Phys. Rev. D 14(2), 632–646. Amati D, Marchesini G, Ciafoloni M & Parisi G 1976 Nuclear Physics B 114, 483–504. Bak P, Tang C & Wiesenfeld K 1988 Physical Review A 38(1), 364–374. Beggs J & Plenz D 2003 J. Neurosci. 23(35), 11167–11177. Buice M A & Cowan J D 2007 Physical Review E 75, 051919. Buice M & Cowan J 2009 Prog. Biophys. Theor. Biol. 99(2,3), 53–86. Burns B D 1951 J. Physiol. 112, 156–175. Doi M 1976 J. Phys. A: Math Gen. 9(9), 1465–1477. Gil L & Sornette D 1996 Physical Review Letters 76(21), 3991–3994. Glauber R 1963 Phys. Rev. Lett. 10(3), 84–86. Lampl I, Reichova I & Ferster D 1999 Neuron 22, 361–374. Self-organized criticality in a neural network 17 Levina A, Herrmann J & Geisel T 2007 Nature Physics 3(12), 857–860. Millman D, Mihalas S, Kirkwood A & Niebur E 2010 Nature Physics 6(10), 801–805. Nauhaus I, Busse L, Carandini M & Ringach D 2009 Nature Neurosci. 12(1), 70–76. Peliti L 1985 J. Physique 46, 1469–1483. Schrodinger E 1926 Naturwissenschaften 14, 664–666. Stevens C 1989 Neural Comp. 1, 473–479. Van Kampen N 1981 Stochastic Processes in Physics and Chemistry North Holland. Vogels T, Sprekeler H, Zenke F, Clopath C & Gerstner W 2011 Science 334(6062), 664–666. Wilson H & Cowan J 1972 Biophys. J. 12, 1–22. Wilson H R & Cowan J D 1973 Kybernetik 13, 55–80.
1801.04623
1
1801
2018-01-15T00:10:29
Sex differences in network controllability as a predictor of executive function in youth
[ "q-bio.NC", "eess.SY" ]
Executive function emerges late in development and displays different developmental trends in males and females. Sex differences in executive function in youth have been linked to vulnerability to psychopathology as well as to behaviors that impinge on health. Yet, the neurobiological basis of these differences is not well understood. Here we test the hypothesis that sex differences in executive function in youth stem from sex differences in the controllability of structural brain networks as they rewire over development. Combining methods from network neuroscience and network control theory, we characterize the network control properties of structural brain networks estimated from diffusion imaging data acquired in males and females in a sample of 882 youth aged 8-22 years. We summarize the control properties of these networks by estimating average and modal controllability, two statistics that probe the ease with which brain areas can drive the network towards easy- versus difficult-to-reach states. We find that females have higher modal controllability in frontal, parietal, and subcortical regions while males have higher average controllability in frontal and subcortical regions. Furthermore, average controllability values in the medial frontal cortex and subcortex, both higher in males, are negatively related to executive function. Finally, we find that average controllability predicts sex-dependent individual differences in activation during an n-back working memory task. Taken together, our findings support the notion that sex differences in the controllability of structural brain networks can partially explain sex differences in executive function. Controllability of structural brain networks also predicts features of task-relevant activation, suggesting the potential for controllability to represent context-specific constraints on network state more generally.
q-bio.NC
q-bio
Sex differences in network controllability as a predictor of executive function in youth Eli J. Cornblath1,2, Evelyn Tang2, Graham L. Baum1,2,3, Tyler M. Moore3, David R. Roalf3, Ruben C. Gur3, Raquel E. Gur3, Fabio Pasqualetti4, Theodore D. Satterthwaite3,*, and Danielle S. Bassett2,5,6,7,* 1Department of Neuroscience, University of Pennsylvania, Philadelphia, PA 19104 USA 2Department of Bioengineering, University of Pennsylvania, Philadelphia, PA 19104 USA 3Department of Psychiatry, University of Pennsylvania, Philadelphia, PA 19104 USA 4Department of Mechanical Engineering, University of California, Riverside, CA 92521 USA 5Department of Electrical and Systems Engineering, University of Pennsylvania, Philadelphia, PA 19104 USA 6Department of Neurology, University of Pennsylvania, Philadelphia, PA 19104 USA 7To whom correspondence should be addressed: [email protected]. *These authors contributed equally. January 16, 2018 8 1 0 2 n a J 5 1 ] . C N o i b - q [ 1 v 3 2 6 4 0 . 1 0 8 1 : v i X r a 1 Abstract Executive function is a quintessential human capacity that emerges late in development and dis- plays different developmental trends in males and females. Sex differences in executive function in youth have been linked to vulnerability to psychopathology as well as to behaviors that impinge on health, wellbeing, and longevity. Yet, the neurobiological basis of these differences is not well understood, in part due to the spatiotemporal complexity inherent in patterns of brain network maturation supporting executive function. Here we test the hypothesis that sex differences in exec- utive function in youth stem from sex differences in the controllability of structural brain networks as they rewire over development. Combining methods from network neuroscience and network con- trol theory, we characterize the network control properties of structural brain networks estimated from diffusion imaging data acquired in males and females in a sample of 882 youth aged 8-22 years. We summarize the control properties of these networks by estimating average and modal control- lability, two statistics that probe the ease with which brain areas can drive the network towards easy- versus difficult-to-reach states. We find that females have higher modal controllability in frontal, parietal, and subcortical regions while males have higher average controllability in frontal and subcortical regions. Furthermore, average controllability values in the medial frontal cortex and subcortex, both higher in males, are negatively related to executive function. Finally, we find that average controllability predicts sex-dependent individual differences in activation during an n-back working memory task. Taken together, our findings support the notion that sex differences in the controllability of structural brain networks can partially explain sex differences in executive function. Controllability of structural brain networks also predicts features of task-relevant activa- tion, suggesting the potential for controllability to represent context-specific constraints on network state more generally. Keywords: network controllability, neurodevelopment, sex differences, executive function, working memory, fMRI BOLD, diffusion tensor imaging Introduction Executive function is necessary for regulation of goal-directed behavior, and encompasses cogni- tive processes including working memory, inhibition, task switching, and performance monitor- ing 1. Deficits in executive function are associated with increased risk taking and associated con- sequences 2,3, and more generally hamper academic and occupational performance 4. Executive deficits frequently lead to personal, social, and professional consequences that accumulate through- out the course of a patient's life 4. Importantly, the normative capacity for executive function rapidly increases during adolescence and differs by sex. Sex differences in the developmental tra- jectory of executive function have been linked to higher rates of impulsivity 5, ADHD diagnosis 6, criminality 7 and substance use 2 in males. Current interventions for disorders of executive function are relatively limited, usually do not consider sex, and rely primarily on psychotherapy and global manipulations via psychopharmacology 8. A basic understanding of sex-related differences in executive function and their implications for the diagnosis and treatment of executive deficits requires an understanding of the normative matura- tion of underlying neural circuitry. Several recent studies highlight the fact that such maturation, and sex-differences in that maturation, span structure 9, anatomical connectivity 10, functional ac- tivity 11,12,13, and functional connectivity 14. Using structural MRI, a recent study 15 reported age- related, non-linear increases in gray matter density with concurrent decreases in cortical thickness. Interestingly, the maturation of these structural features was markedly different between the sexes, with females showing higher gray matter density globally and higher cortical thickness in frontal and insular regions 15. Several older structural studies 16 have also reported linear increases in fron- toparietal white matter density throughout adolescence. Using diffusion-weighted MRI, another study found significantly greater within-hemisphere connectivity in males and greater between- hemisphere connectivity in females 17. Sex differences have also been identified in the clustered (or modular) structure in patterns of functional connectivity estimated from resting state fMRI data: males display higher between-module connectivity while females display higher within-module con- nectivity 18, and these differences were shown to predict individual differences in executive func- tion 18. Although these descriptive studies have provided important insights, it remains difficult to specify in a mechanistic sense how executive function might arise from such complex, multimodal patterns of brain maturation in a sex-dependent manner. We address this challenge by positing that sex differences in executive function in youth stem from sex differences in the controllability of structural brain networks as they rewire over development. This notion intuitively bridges the control of behavior (executive function) with the control of brain dynamics (network controllability). Specifically, we capitalize on recent advances in network control theory 19,20, an emerging branch of theoretical physics and systems engineering that builds on early efforts in control theory 21,22 to offer a mechanistic model of how key nodes, or control points, can exert disproportionate influence over system function 19. Control points are identified with metrics that assess the ability of specific nodes to alter a system's state, based on the underlying network topology 20 (Fig. 1). Specifically, the metric of average controllability reflects the average energy input required at a node to move the system from some initial state to all possible states. In contrast, the metric of modal controllability reflects the ease of transitioning the system from some initial state to a difficult-to-reach state. Prior work has demonstrated the utility of network control theory in understanding basic brain architecture and function 23,24,25 across spatial scales 26 and 3 species 27, posited its relation to cognition 28,29, and outlined its developmental course 29. Here, we tested the hypothesis that developmental sex differences in network control underlie sex differences in executive functioning. Specifically, we predicted that (i) network controllability differs by sex, (ii) network controllability changes with age differently in males and females, (iii) sex differ- ences in network controllability predict executive function, and (iv) network controllability predicts the activation of brain regions during a working memory task demanding executive function. To test these hypotheses, we constructed structural brain networks from diffusion tensor imaging data acquired in 882 healthy youth, ages 8-22 years, in the Philadelphia Neurodevelopmental Cohort (PNC) 30. Each brain network was comprised of 234 anatomically defined brain regions 31 connected by white matter tracts estimated from diffusion tractography. We show that regional controllability is a significant mediator of the relationship between sex and executive function, and that it predicts the magnitude of fMRI BOLD signal on an n-back working memory task. As described in detail below, our results suggest that sex differences in the controllability of structural brain networks predict executive function and the activity profiles that support that function. Materials and Methods Participants Diffusion tensor imaging (DTI) data were obtained from youth who participated in a large community- based study of brain development, now known as the Philadelphia Neurodevelopmental Cohort (PNC) 30. Here we study 882 out of a total of 1601 subjects between the ages of 8 and 22 years (mean age = 15.06, SD = 3.15, 389 males, 493 females). Due to lack of complete diffusion scans (n = 224) and incidental findings (n = 20), data from 244 participants was deemed unusable. The remaining 1357 participants underwent a rigorous manual and automated quality assurance proto- col for DTI datasets 32, eliminating an additional 147 subjects with poor data quality. A subset of 93 of the remaining 1210 participants were excluded for low quality or incomplete FreeSurfer re- construction of T1-weighted images. Further, 235 of the remaining 1117 participants were excluded for one or more of the following reasons: gross radiological abnormalities distorting brain anatomy, medical history that might impact brain function, history of inpatient psychiatric hospitalization, use of psychotropic medication at the time of imaging, or high levels of in-scanner head motion during the DTI scan, as defined by a mean relative displacement between non-weighted volumes of greater than 2 mm. These exclusions left us with a final sample of n = 882 subjects 10,29 between the ages of 8 and 22 years (mean age = 15.06, SD = 3.15, 389 males, 493 females). Cognitive Phenotyping Cognition was measured outside of the scanner using the Penn Computerized Neurocognitive Bat- tery (CNB) 33,34. Briefly, the 1-hour CNB was administered to all participants, and consisted of 14 tests that evaluated a broad range of cognitive functions. Twelve of the tests measure both accuracy and speed, while two of the tests (motor and sensorimotor) measure only speed. Overall cognitive performance was summarized as the average z-transformed accuracy and speed (as well as the difference between accuracy and median response time: efficiency) scores across all tests administered (as described in Moore et al. 35 for complete details). Factor scores are described in 4 Moore et al. 35; here, we used the factor score for executive efficiency from a best-fitting four-factor solution comprising tests from the executive function domain (attention, abstraction and working memory). The tests contributing to the executive efficiency score include the Penn Continuous Performance Test, the Letter N-Back task, and (weakly) the Penn Verbal Reasoning Test 35,36,37 (analogical reasoning). In the present study, we use this factor score for executive efficiency as our primary measure of executive function, hereafter referred to as "executive function". Imaging Data Acquisition MRI data were acquired on a 3 Tesla Siemens Tim Trio whole-body scanner and 32-channel head coil at the Hospital of the University of Pennsylvania. DTI scans were acquired via a twice-refocused spin-echo (TRSE) single-shot echo-planar imaging (EPI) sequence (TR = 8100ms, TE = 82ms, FOV = 240mm2/240mm2; Matrix = RL:128/AP:128/Slices:70, in-plane resolution (x and y) 1.875 mm2; slice thickness = 2mm, gap = 0; flip angle = 90/180/180 degrees, volumes = 71, GRAPPA factor = 3, bandwidth = 2170 Hz/pixel, PE direction = AP). This sequence utilizes a four-lobed diffusion encoding gradient scheme combined with a 90-180-180 spin-echo sequence designed to minimize eddy-current artifacts. The complete sequence consisted of 64 diffusion-weighted directions with b = 1000s/mm2 and 7 interspersed scans where b = 0s/mm2. Total scan time was approximately 11 min. The imaging volume was prescribed in axial orientation covering the entire cerebrum with the topmost slice just superior to the apex of the brain. In addition to the DTI scan, a map of the main magnetic field (i.e., B0) was derived from a double-echo, gradient-recalled echo (GRE) sequence, allowing us to estimate field distortions in each dataset. Prior to DTI acquisition, a 5-minute magnetization-prepared, rapid acquisition gradient-echo T1-weighted (MPRAGE) image (TR 1810 ms, TE 3.51 ms, FOV 180 ×240 mm, matrix 256 × 192, effective voxel resolution of 1 × 1 × 1 mm) was acquired. This high-resolution structural image was used for tissue segmentation and parcellating gray matter into anatomically defined regions in native space. Rigorous manual and automated quality-assurance protocols for the T1-weighted structural imaging data were performed and cleared for the 882 subjects consid- ered here 38. Subsequently, all structural images were processed using FreeSurfer (version 5.3) 39. FreeSurfer reconstructions underwent rigorous quality-assurance protocols 38,40. The T1 image was parcellated into 234 regions by FreeSurfer according to the Lausanne Atlas 41. We define the sub- cortex of this 234-region parcellation to be comprised of the left and right hemispheric counterparts of the thalamus proper, caudate, putamen, pallidum, nucleus accumbens area, hippocampus and amygdala, while excluding the brainstem (14 regions). We define the cortex of this 234-region parcellation to be comprised of the remaining regions (219 regions). fMRI BOLD data were acquired as subjects completed a version of the n-back task using fractal images 42. See the supplementary methods of ref. 43 for details regarding task presentation and structure. Functional images were obtained using a whole-brain, single-shot, multislice, gradient- echo echoplanar sequence (231 volumes; TR = 3000 ms; TE = 32 ms; flip angle = 90 degrees; FOV = 192× 192 mm; matrix = 64× 64; slices = 46; slice thickness = 3 mm; slice gap = 0 mm; effective voxel resolution = 3.0 × 3.0 × 3.0 mm). 5 Imaging Data Preprocessing All DTI datasets were subject to a rigorous manual quality assessment procedure involving visual inspection of all 71 volumes 32. Each volume was evaluated for the presence of artifact, with the total number of volumes impacted summed over the series. This scoring was based on previous work describing the impact of removing image volumes when estimating the diffusion tensor 44,45. Data was considered "Poor" if more than 14 (20%) volumes contained artifact, "Good" if it contained 1-14 volumes with artifact, and "Excellent" if no visible artifacts were detected in any volumes. All 882 subjects included in the present study had diffusion datasets identified as "Good" or "Excellent," and had less than 2mm mean relative displacement between interspersed b = 0 volumes. As described below, even after this rigorous quality assurance, motion was included as a covariate in all analyses. The skull was removed for each subject by registering a binary mask of a standard fractional anisotropy (FA) map (FMRIB58 FA) to each subject's DTI image using an affine transformation 46. Eddy currents and subject motion were estimated and corrected using the FSL eddy tool 47. Diffu- sion gradient vectors were then rotated to adjust for subject motion estimated by eddy. After the field map was estimated, distortion correction was applied to DTI data using FSL's FUGUE 48. BOLD time series were processed as described in 49,43. Briefly, FSL 5 48 was used to analyze time se- ries data from three condition blocks (0-back, 1-back and 2-back), with the primary contrast being 2-back > 0-back. BOLD images were co-registered to the T1 image using boundary-based regis- tration 50 with integrated distortion correction as implemented in FSL. Generalized linear model (GLM) beta weights were averaged across all voxels in each parcel of the 234-node Lausanne atlas. In our assessment of n-back performance-related activation, we use the difference in GLM beta weights between the 2-back and 0-back condition. For all analyses of fMRI data, we excluded 223 subjects with incomplete data or excessive head motion (mean relative displacement > 0.5 mm or maximum displacement > 6 mm), leaving n = 659 remaining. Structural Network Estimation Structural connectivity was estimated from DTI data in order to generate the adjacency matrix representing the pattern of white matter tracts between large-scale brain areas. DSI Studio was used to estimate the diffusion tensor and perform deterministic whole-brain fiber tracking with a modified FACT algorithm that used exactly 1,000,000 streamlines per subject after removing all streamline with length < 10 mm 28. To extend regions into white matter, parcels defined using the Lausanne atlas were dilated by 4 mm 28,29 and registered to the first non-weighted (b = 0) volume using an affine transform 28,29. The number of streamlines connecting each node of the 234 region parcel end-to-end was used to define the edge weights aij of the adjacency matrix A. Network Controllability We represent the streamline-weighted structural network estimated from diffusion tractography as the graph G = (V,E), where V and E are the vertex and edge sets, respectively. Let aij be the weight associated with the edge (i, j) ∈ E, and define the weighted adjacency matrix of G as A = [aij], where aij = 0 whenever (i, j) /∈ E. We associate a real value with each node to generate a vector 6 describing the network state, and we define the map x : N≥0 ← RN to describe the dynamics of the network state over time. It is worth noting that this method assumes that the number of streamlines is proportional to the strength of structural connectivity with regards to propagation of activity between nodes according to a specified model of dynamics. Here we employ a simplified noise-free linear discrete-time and time-invariant model of such dynamics: x(t + 1) = Ax(t) + BKuK(t), (1) where x describes the state (i.e. voltage, firing rate, BOLD signal) of brain regions over time. Thus, the state vector x has length N , where N is the number of brain regions in the connectome parcellation, and the value of xi describes the brain activity state of that region. The matrix A is symmetric, with the diagonal elements satisfying Aii = 0. Prior to calculating controllability values, we divide A by 1 + ξ0(A), where ξ0(A) is the largest eigenvalue of A. The input matrix BK identifies the control point K in the brain, where K = k1, ..., km and (2) and ei denotes the i-th canonical vector of dimension N . The input uK : R≥0 → RM denotes the control strategy. BK = [ek1 ··· ekm], To study the dynamics by which the activity of one brain region influences structurally connected regions, we apply the control theoretic notion of controllability to our dynamical model. Classic results in control theory ensure that controllability of the network, x(t + 1) = Ax(t) + BKuK(t), from the set of network nodes K is equivalent to the controllability Gramian WK being invertible, where ∞(cid:88) τ =0 WK = Aτ BKBTKAτ . (3) We calculate WK, or rather Wi, with BK set equal to one canonical vector ei and repeat this process for all N nodes 29,28. Although it is well known that the activity of several brain regions and neuronal ensembles are related via non-linear dynamics, it has been shown that a linear approximation can explain features of the resting state fMRI BOLD signal 51; this suggests that a linear approximation can effectively capture the controllability properties of the original non-linear dynamics. Controllability Metrics Following ref. 29,28, controllability metrics for structural brain networks were calculated for two different control strategies which describe the ability to change x(t) in a particular fashion 20. 7 Average controllability describes the ease of transition to energetically similar states, while modal controllability describes the ease of transition to difficult-to-reach states 20. Average controllability of a network equals the average input energy applied to a set of control nodes required to reach all possible target states. It is known that average input energy is proportional cannot be accurately computed even at more coarse connectome parcellations. In addition to its to Trace(W−1K ), the trace of the inverse of the controllability Gramian. Instead, as in ref. 29,28, we use Trace(WK) as a measure of average controllability because (i), Trace(WK) and Trace(W−1K ) are related via inverse proportionality, and (ii), Trace(W−1K ) tends to be very ill-conditioned and relationship with Trace(W−1K ), Trace(WK) describes the energy of the network impulse response, or, equivalently, the network H2 norm 21,22. In summary, to compute the average controllability value for node i in A, we compute the Trace(WK) when node i is the only control node (i.e. BK = ei). Modal controllability refers to the ability of a node to control the evolutionary modes of a dynamical network, and is most interpretable when used to identify states that are poorly controllable given BK. To calculate modal controllability, one must first obtain the eigenvector matrix V = [vij] of the adjacency matrix A. If vij is small, then the j-th evolutionary mode of the input-independent form of Eq. (1), x(t) = Ax(t), is poorly controllable from node i. According to prior work 20, we ij as a scaled measure of the modal controllability of each of the N define φi =(cid:80)N modes ξ0(A), ..., ξN−1(A) from node i. j=1(1 − ξ2 j (A))v2 Finally, we calculate the mean of the regional controllability values either across the whole brain (Fig. A.1) or within the cortex and subcortex separately (Figs. 3, 4, 5, 6). For each subject, we define the mean modal controllability to be the sum of the values of φi for each node within A, divided by the number of regions. Similarly, we define the mean average controllability to be the sum of the values of Trace(WK) for each node within A, divided by the number of regions. Network Synchronizability While the metrics of average and modal controllability provide complementary views on the diverse dynamics that a brain network can display, it is also useful to study contrasting metrics that probe the susceptibility of the network to be constrained within a narrower range of dynamics. Previous work 29 has demonstrated that a useful contrasting metric to controllability is synchronizability, which intuitively measures the susceptibility of a network to remain in a single synchronous state s(t), i.e. x1 = ··· = xn(t + 1) = s(t). The master stability function (MSF) allows one to analyze (cid:80) the stability of this synchronous state without fully characterizing the properties of each feature of the system 52. Within this framework, linear stability depends on the positive eigenvalues λi, i = 1, ..., N − 1 of the Laplacian matrix L defined by Lij = δij k Aik − Aij, where δij is the Kronecker delta function. The condition for stability depends on the shape of the MSF, i.e. the stability of the synchronized state to linear perturbations holds when the MSF is negative for all non-zero eigenvalues of L. This condition is more likely for a smaller range of eigenvalues, hence we can use the normalized spread of these eigenvalues to quantify network synchronizability as 8 (cid:80)N−1 d2(N − 1) i=1 λi − ¯λ2 1 σ2 = , where ¯λ := 1 N − 1 N−1(cid:88) i=1 λi (4) (cid:80) (cid:80) and d := 1 j(cid:54)=i Aij, the average coupling strength per node, which normalizes for the overall N network strength. Using this definition, we calculated the synchronizability of A for each subject. i Linear Regression of Network Control Metrics on Sex For all analyses of network control metrics, we examined the effect of sex while controlling for age, total brain volume (segmented brain volume, as defined by FreeSurfer BrainSegVol metric), handedness, and motion during the diffusion scan. We used multiple ordinary least squares (OLS) linear regression with the lm() command in R or the MATLAB regress function to fit the following general equation: C = 1 + βaa + βvv + βhh + βmdmd + βss , (5) where C is the controllability statistic (either φi for modal controllability or A for average con- trollability), a is age, v is total intracranial volume, md is the mean framewise displacement as a summary measure of in-scanner head motion during the diffusion imaging sequence, h is handed- ness, and s is sex. We then used the R package visreg to calculate 95% confidence intervals around fitted lines and generate partial residuals. Furthermore, when using the multiple OLS regression for the node-level analyses of controllability, we applied a false discovery rate (FDR) correction 53 (q < 0.05) to control for Type I error due to multiple testing. Non-Linear Fits of Executive Function and Network Metrics Because age and executive function are nonlinearly related, we used a general additive model 54,38 (GAM) to assess the relationship between executive efficiency and age using the mgcv package in R (Fig. 1b). A GAM is a generalized linear model in which the linear predictor is defined by unknown smooth functions of predictor variables. In this case, we use penalized regression splines for age, with conventional parametric components for the remaining predictors. Following ref. 29, we used non-linear models to examine the relationships between synchronizability, average controllability, and modal controllability (Fig. 5). We generated estimates of parameters for models of the form y = a+b exp(cx) via non-linear least squares using the nls() function in R. To compare these fits between males and females, we performed an analysis of variance (ANOVA). For example, when considering sex differences in the nonlinear relation between modal controllability and average controllability, we examined the expression: φi = (a + αs) + (b + βs) exp((c + γs)A) , (6) where φi is modal controllability, A is average controllability, and s is sex, coded as 0 for males and 1 for females, so that α, β and γ = 0 for males and the equation is reduced to the base form. 9 The predicted values for males or females obtained from the full model were used to generate the curves in Fig. 5. Mediation Testing After identifying associations between sex, controllability, and executive function, we asked whether controllability is a statistically significant mediator of the relationship between sex and executive function. To test for a formal mediation, we performed non-parametric bootstrapping causal me- diation analysis using the R function mediate() from the package mediation 55. It is worth noting that we cannot estimate causality in cross-sectional association data. The Average Causal Me- diation Effect (ACME) quantifies the relationship between mediator and outcome independent of treatment, while the Average Direct Effect (ADE) quantifies the relationship between treatment and outcome, independent of the mediator 55. Linear Regression of BOLD Data on Network Control Metrics Finally, and in keeping with our linear systems framework, we sought to test whether controllability would predict brain state (x(t)), as defined by BOLD activation during a task demanding executive function. In these analyses (Fig. 8), we used the residuals from Eq. 5 for controllability. For activation, we used the residuals from the following equation: A = 1 + βaa + βhh + βmnmn + βss , (7) where A is 2-back minus 0-back activation (hereafter referred to as "activation"), a is age, h is handedness, mn is the mean framewise displacement during the n-back scan, and s is sex. Because we had no prior knowledge about where controllability might predict activation, we fit linear models between controllability and activation at every possible pair of nodes, i.e. we performed 234 × 234 regressions of controllability residuals at each node with activation residuals at each node. The 234× 234 matrices of p-values for the slope of controllability on activation were FDR corrected (q < 0.05) separately for average controllability and for modal controllability. To identify regions where activation was associated with executive function, we followed previous work 49,43 by examining d(cid:48), a composite measure of n-back task performance which takes both correct responses and false positives into account to separate performance from response bias. Results Sex differences in executive function Our analysis of sex differences in the development of executive function and its neurobiological underpinnings began with a sex-stratified comparison of mean performance on executive efficiency, a cross-contextual summary measurement of executive function 35. Consistent with prior results from the PNC data 18,30, the executive efficiency subscore of the PCNB was significantly higher in females (full model r2 = 0.52, df = 873, p = 0.0015; Fig. 2a) with no interaction between age and sex. Furthermore, executive efficiency increased with age in a non-linear fashion (r2 = 0.52, df = 2, p < 10−15; Fig. 2a). This result suggests that development of executive function occurs via a similar 10 course for males and females, but that females display higher scores in this cognitive domain for all ages studied. After performing this sex-stratified analysis of the developmental course of the executive efficiency score, we next turned to a consideration of its neurobiological underpinnings (Fig. 1a). Sex differences in regional network controllability Our general hypothesis was that sex differences in executive function in youth stem from sex differences in the controllability of structural brain networks as they rewire over development, a notion that bridges the control of behavior (executive function) with the control of brain dynamics (network controllability). To test this hypothesis, we examined the anatomical distribution of control points in the structural brain networks of males and females separately. We ranked the mean average controllability value at each region for males and females separately, which revealed that the distribution of controller strength is virtually identical between male and females (r = 0.99; Fig. 2b). This result suggests that there is no sex difference in the spatial distribution of controllers when classified by their relative magnitudes. We next investigated whether the actual (rather than ranked) regional controllability estimates dif- fered by sex, and we addressed this question first by considering the cortex and subcortex separately. Specifically, we computed mean controllability values for each subject across 219 cortical regions In the cortex, mean average controllability (r2 = 0.082, df = 876, and 14 subcortical regions. p = 0.018; Fig. 3a) and mean modal controllability (r2 = 0.13, df = 876, p = 0.018; Fig. 3a) are higher in females. Mean modal controllability in the subcortex is also higher in females (r2 = 0.047, df = 876, p < 10−4; Fig. 3b), whereas mean average controllability in the subcortex is higher in males (r2 = 0.036, df = 876, p = 0.034; Fig. 3b). Only with average controllability in the subcortex did males have higher controllability than females; this result suggests that the connectivity profile of subcortical regions may contribute to sex differences in functional brain dynamics. In a finer-grained analysis, we investigated whether the controllability of single regions differed by sex. Modal and average controllability were separately examined at each region, while accounting for age, total brain volume, handedness, and mean in-scanner head motion as model covariates. We found that average and modal controllability differed by sex at a subset of network nodes after FDR correction (q < 0.05) for multiple comparisons. Nodes with average controllability values that differed by sex were almost all (4/5) higher in males and located in the frontal lobe or subcortex (Fig. 3c-d). Conversely, nodes with modal controllability values that differed by sex were all (18/18) higher in females and located in frontoparietal and subcortical systems. These results suggest that controller strength differs between males and females on a region-specific and control strategy-specific basis. Development of network controllability across the sexes We next turned to assessing whether the developmental arc of network controllability differed by sex. We found that controllability in the cortex and subcortex tended to increase with age, with In the cortex (r2 = 0.13, df = 876, the exception of average controllability in the subcortex. p = 9.9 × 10−10; Fig. 4b) and subcortex (r2 = 0.047, df = 876, p = 1.0 × 10−5; Fig. 4d), mean modal controllability increases with age for males and females. Mean cortical average controllability also increases with age (r2 = 0.082, df = 876, p = 1.9 × 10−15; Fig. 4b). Interestingly, there was a 11 significant age-by-sex interaction only with subcortical average controllability, such that the slope was positive for males and negative for females (Fig. 4c). However, we also found that subcortical average controllability remains stable throughout development for both males (βage = 0.47, t = 1.29, p = 0.20, df = 875) and females ((βage + βage∗sex) = −0.65, p = 0.052, t = −1.94, df = 875) 56. Taken together, these results suggest that controllability changes with age similarly for males and females, but that average controllability in the subcortex changes with age differently than modal controllability or than cortical average controllability. With the knowledge that controllability has a sex-independent relationship with age, we were interested in testing the hypothesis that sex influences the relationship between different types of controllability in the developing brain. Following ref. 29, we fit the relationships between average and modal controllability with the exponential function y = a + b exp(cx) separately for each sex (Eq. 6). Our results showed that males and females did not have a statistically different relationship between average and modal controllability averaged across the whole brain (p = 0.16, df = 3; Fig. 5a,e). In the cortex alone, average and modal controllability followed an increasing exponential form (Fig. 5b,f), similar to that of the whole brain. In contrast, in the subcortex alone average and modal controllability followed a decreasing exponential form (Fig. 5d,h). When sex was included in the model, fits improved significantly (cortex: p = 2.1 × 10−8, df = 3, Fig. 5b,f; subcortex: p = 6.3 × 10−9, df = 3, Fig. 5c,g), suggesting that these regions may contain sex-dependent differences in structural connections important for controlling brain network state. Next, we performed a specificity analysis to determine whether our results could be further con- firmed by sex differences in a contrasting metric – synchronizability – that probed the susceptibility of the network to be constrained within a narrow (rather than broad) range of dynamics. We found that synchronizability did not differ between males and females (A.1b) and decreased with age in a sex-independent fashion (A.1d). Moreover, the exponential relationship between synchronizability and average controllability did not differ by sex (p = 0.33, df = 3; Fig. 5d,h), confirming the local (Fig. 3c-d) rather than global (Fig 2b) differences in metrics of network control and dynamics. Sex differences in network controllability predict individual differences in exec- utive function While sex differences in network controllability are of interest in understanding the structural drivers of brain dynamics, their impact on behavior requires a link to cognition. Here we examine the relation between executive function and network controllability across cortex and subcortex separately, and then at individual brain regions. When considering mean cortical and subcortical controllability and executive function we found that neither average (r2 = 0.46, df = 873, p = 0.96; Fig. 6a) nor modal controllability (r2 = 0.46, df = 873, p = 0.25; Fig. 6b) in the cortex were asso- ciated with performance. A different trend was apparent in the subcortex: average controllability was negatively correlated with performance (r2 = 0.47, df = 873, p = 5.4 × 10−4; Fig. 6c), while modal controllability was positively correlated with performance (r2 = 0.47, df = 873, p = 0.024; Fig. 6d). Next, we considered the 21 brain areas that we had previously found to display sex differences in controllability values (5 nodes for average controllability and 18 nodes for modal controllability, with 2 nodes overlapping). Nodes with significantly higher average controllability in males (right medial orbitofrontal cortex and bilateral caudate nuclei) had average controllability values that 12 were negatively related to executive function (Fig. 6e; FDR corrected, q < 0.05). Interestingly, modal controllability at any individual node was not significantly associated with executive function after FDR correction; however, two nodes with uncorrected p < 0.05 (right caudate nucleus and left inferior parietal lobe) had modal controllability values that were higher in females and positively associated with executive function. This result suggests that modal controllability of one particular node is not as strong of a predictor of cognitive performance as mean modal controllability over the subcortex as a whole. One parsimonious explanation for the results reported up to this point is that controllability medi- ates the relationship between sex and executive function: specifically, subcortical average controlla- bility is higher in males, and increasing subcortical average controllability is associated with poorer performance on an assessment of executive function. We explicitly tested for such a mediation and found that, indeed, average controllability at the right medial orbitofrontal cortex, right superior frontal cortex, right postcentral gyrus and the bilateral caudate nuclei were statistically significant mediators of the relationship between sex and executive function (Fig. 7a). Similarly, subcortical modal controllability was higher in females, and increasing subcortical modal controllability was associated with better executive function. However, consistent with the fact that no individual node had modal controllability values associated with executive function, modal controllability was not found to be a mediator of the relationship between sex and executive function. Taken together, these results suggest that sex differences in average controllability may be a predictive biomarker for sex differences in executive function. Sex-dependent relationships between n-back task activation magnitudes and con- trollability Describing the relationship between sex, regional controllability, and cognitive performance provides us with a better understanding of the importance of structural brain networks in sex differences in executive function. The final aspect of our hypothesis pertains to whether network control theory can be used to explain how differences in brain network structure produce divergent patterns of brain activity that underpin executive function. We hypothesized that regional controllability values would predict regional n-back task activation magnitudes, and that associations between controllability and activation would differ by sex. To obtain a reference point for activation profiles associated with strong executive function, we separately regressed activation at each node on d(cid:48) 49. This analysis identified 113 nodes for which activation was associated with successful task performance (Fig. 8a). Note that in contrast to the use of a general factor score for executive efficiency in the previous analyses, here we considered the d(cid:48) measure to summarize performance on the n-back task only 57 so that the performance measure is directly related to the context in which activation values are measured. Interestingly, sex was not associated with activation at any brain region (FDR corrected, q < 0.05, data not shown). After identifying regions at which activation was associated with executive function, we sought to relate these activation values to controllability metrics of structural brain networks. We found a complex pattern of both positive and negative relationships between average controllability at 8 nodes with activation at 11 nodes (Fig. 8b). Average controllability at the left caudal middle frontal lobe was associated with activation at 7 different regions, which cluster together in a symmetric fashion in superior parietal and lateral occipital cortex (Fig. 8c). Modal controllability was not 13 significantly associated with activation. Next, we tested our hypothesis that the relationship between regional controllability and activation depends on sex. Specifically, we performed the same analysis on males and females, separately. Consistent with the pooled analysis, the profile of controllability-activation associations overlapped significantly between the sexes (Fig. 8d, Fig. A.4a). While there was no overlap in the precise regions involved in these sex-split analyses, there were gross anatomical similarities between males and females. In both sexes, controllability at regions near the temporal poles was associated with activation at medial frontal regions. We observed no significant interaction between controllability and sex in predicting activation (Fig. A.4b), although the directionality of the interaction terms was consistent with the observed sex differences (Fig. A.4a-b). These results support the notion that both similarities and differences in structure-function relationships exist across the sexes. Weighted degree does not explain controllability trends Weighted degree has previously been shown to correlate well with controllability metrics 28. We performed several specificity analyses to determine whether our results could be trivially explained by sex-differences in the weighted degree of nodes in the network. We identified 13 nodes where weighted degree was associated with sex (Fig. A.3a), and found that only 5/21 unique nodes with sex-associated controllability values also had a significant association between weighted degree and sex. Among those 5 unique nodes, only average controllability at the left caudate nucleus and the right post central gyrus had the same directionality of association with sex as the association with weighted degree (i.e. higher in males and females, respectively). When weighted degree was averaged across the anatomical cortex or subcortex, there were no significant associations between sex and cortical (r2 = 0.090, df = 876, p = 0.73; Fig. A.2a) or subcortical (r2 = 0.077, df = 876, p = 0.13; Fig. A.2c) weighted degree. These findings suggest that despite the known associations between controllability and weighted degree 28, weighted degree does not explain the observed sex differences in controllability. Next, we tested whether developmental trends of weighted degree paralleled those of controllability metrics. We found that when averaged across the cortex or subcortex, weighted degree is positively associated with average controllability and negatively associated with modal controllability. How- ever, mean weighted degree increased with age in the cortex (r2 = 0.090, df = 876, p = 0.0034; Fig. A.2b), but decreased with age in the subcortex (r2 = 0.077, df = 876, p = 0.011; Fig. A.2d). Neither of these two trends (Fig. A.2b,d) were simple mirrors of the relationships between age and regional average or modal controllability (Fig. 4), suggesting that developmental trends of controllability reflect a more complex phenomenon than developmental trends of weighted de- gree. Furthermore, at nodes with sex-associated weighted degree or sex-associated controllability, weighted degree did not significantly mediate the relationship between sex and executive function (Fig. A.6). These findings suggest that weighted degree does not explain developmental trends of controllability, nor does it mediate the relationship between sex and executive function. Finally, we assessed whether the associations between controllability and activation could be simply explained by the weighted degree of nodes in the network. We found no significant associations between activation and weighted degree in a pooled analysis. Moreover, including each subject's global mean weighted degree in the model did not change the associations between average con- trollability and activation (data not shown). While there were significant associations between 14 activation and weighted degree for males alone (Fig. A.5), weighted degree-activation profiles did not overlap well with average controllability-activation profiles. These results suggest that weighted degree does not explain the relation between controllability and activation. Discussion Our study demonstrates that network control theory can be used to explain how differences in brain network structure produce divergent patterns of brain activity that underpin executive func- tion and its differences across males and females. We first showed that the relative locations of controllers by strength did not differ between males and females, suggesting that the overall struc- ture of control points is similar across sex. Then, we showed that controllability covaried with age in a sex-independent fashion in both the cortex and subcortex. While global and developmental sex differences appear minimal, local sex differences in controllability exist and average controllability at those regions predicts executive function. Notably, average controllability significantly mediated the association between sex and executive function, such that subcortical average controllability was higher in males and is negatively associated with executive function. Crucially, consistent with predictions from network control theory, we also observed sex-dependent associations between nodal average controllability and n-back task BOLD activation magnitudes, demonstrating that differ- ences in brain network structure produce divergent patterns of brain activity supporting executive function. Implications for cognitive and clinical neuroscience Executive function is generally lower in males, which is reflected in higher rates of impulsivity 5, ADHD diagnoses 6, criminality 7, and substance use 2. While it is known that sex differences exist in the prevalence of disorders of executive function and the putative neural circuitry involved 58,59, it is unclear whether the pathophysiology of such disorders is sex-specific or sex-independent. Our study significantly extends the boundaries of knowledge in demonstrating neurophysiological markers of sex differences in executive function, and in couching such markers within a general network control theory of brain function. The present work also provides important groundwork for clinical therapy. The delivery of psychi- atric care is shifting towards personalized, targeted interventions. This shift can be supported by the methodology and computational tools of network neuroscience, where accurate models of brain structure and dynamics can be constructed for single individuals 60,61. Such subject-, age-, and sex-specific models of brain structure and function can directly inform neuromodulatory therapies such as transcranial magnetic stimulation, by offering predictions about how stimulation will affect both the area being stimulated, and other areas connected to it, thereby producing a complex spatiotemporal influence on brain state. Specifically, assuming a model of brain dynamics allows us to determine which nodes could most easily drive transitions in the state of brain activity via some stimulatory input. The network control theory that we use here to study the internal modulation of brain state (via undertaking a task requiring executive function) also makes explicit predictions about the external modulation of brain state (via stimulation or neurofeedback) 62. Indeed, aver- age controllability values were associated with sex-dependent, symmetric patterns of brain activity (Fig. 8c-d). Our results suggest that average controllability values contain useful information about 15 brain state, beyond what can be predicted from a simple streamline-weighted adjacency matrix. These observations are the first steps towards a characterization of brain dynamics that would allow clinicians to predict the impact of stimulation given a subject-, age-, and sex-specific network architecture, thereby producing predictable changes in brain state. An understanding of the relation between network controllability, sex, and executive function could provide important context for the study and diagnosis of neurological disease and psychiatric dis- orders whose prevalence may differ by sex and whose presentation includes alterations in executive function. In our study, the most robust associations between controllability, sex, and executive function are found in the subcortex and the prefrontal cortex. It is widely known that the pre- frontal cortex is important for behavioral planning and working memory 63,64, two key components of executive function. The subcortex alone is less commonly associated with executive function, but there is a strong precedent for the requirement of circuitry between dorsolateral prefrontal cortex (DLPFC) and subcortical structures for executive function 65,66,67,68. Vascular 69 or neu- rodegenerative 68 (i.e. Huntington's disease, Parkinson's disease) lesions to the subcortex give rise to executive dysfunction, similar to what is observed with DLPFC lesions. Moreover, ADHD, a disorder characterized by impulsivity and executive dysfunction, is associated with decreased stri- atal dopamine transporter expression 58, aberrant striatal fMRI BOLD activation 70, and decreased striatal volume 71,72. Sex differences in striatal volume 73 and striatal dopamine receptor binding 59 have also been reported. In light of these results, our study provides an important account of sex differences in subcortical connectivity from the dynamical perspective of network control theory. Our results suggest that the connectivity profile of subcortical structures is related to sex and is important for executive function, supporting the notion that disorders of executive function arise in part from altered structural connectivity. Methodological Considerations Several methodological considerations are pertinent to this work. First, we use a time-invariant, linear model of brain dynamics because its network control properties have been well characterized mathematically; however, it is known that the brain is highly non-linear 74. Yet, the associations between controllability and brain activity that we uncover here suggest that a simple linear model is sufficient to capture some aspect of the underlying brain dynamics. Second, axons transmit information in a unidirectional fashion, but diffusion imaging and associated tractography tools cannot elucidate the directionality of large axonal fiber bundles. Thus, we construct a symmetric adjacency matrix A, assuming bidirectional influence between network nodes, and all interpreta- tions of regional controllability depend on the realism of that assumption. Third, the eigenvalues of a symmetric matrix can only ever be real and thus the system will not oscillate, as neural systems are known to do. Finally, streamline counts and network density vary between scans 75 and with different tract reconstruction methods 76, which can complicate interpretation. Nevertheless, nu- merous DTI-based studies of neuropsychiatric illness have described differences in structural brain networks consistent with clinical and neuroscientific priors 77,78, suggesting that DTI is capable of capturing subtle but meaningful variation in structural connectivity. It is also important to acknowledge that our analysis of sex differences focuses on sex assigned at birth, which we refer to as "sex" throughout this paper. Our data does not include endocrinological measurements relevant to sex, nor does it include any psychosocial assessment of gender identity. 16 Thus, we were not able to control for or assess gender-based differences in brain structure. Fur- thermore, while there exist two distinct classes of human genitalia, this fact does not imply that brains are also sexually dimorphic 79. Both "male" and "female" features exist in both male and female brains, though some features are more common in one sex than the other 79,80. As a result, there may be more meaningful variance in neurologic phenotypes within each sex than between sexes. The divergent controllability-activation profiles in Fig. 8d may be in part due to this fact. Nevertheless, biological sex is an easily measured variable and identifying correlates of sex makes it a useful biomarker. In the future, identification of additional covariates might help uncover a more ubiquitous reason for sex-associated brain features 80. Conclusion and Future Directions First and foremost, our analysis of sex differences in structural brain networks showed that males and females are highly similar from the perspective of network controllability. The organization of relative controller strength was almost identical between males and females, and sex was not significantly associated with controllability values at most brain regions. However, the differences in average controllability between males and females predicted differences in cognitive performance and effects were most robust in subcortical and frontal regions. Given that BOLD signal is associated with network controllability, an interesting future study might use time-invariant, linear dynamics to predict changes in brain activity after stimulation of regions with high versus low controllability. Our results pave the way for a future study to use sex and age as features that may delineate the likely responses to brain stimulation a priori, obviating the need for pre-stimulation diffusion image acquisition and processing. Such an approach would allow the delivery of personalized psychiatric care at low cost with minimal requirements for imaging equipment and computational power. Acknowledgments This work was supported by an administrative supplement to NIH R21-M MH-106799 (Satterth- waite/Bassett MPI). DSB also acknowledges support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Research Office through contract number W911NF-14-1-0679, the Army Research Laboratory through contract number W911NF-10-2-0022, the National Institute of Health (2-R01-DC-009209-11, 1R01HD086888-01, R01-MH107235, R01- MH107703, R01MH109520, 1R01NS099348 and R21-M MH-106799), the Office of Naval Research, and the National Science Foundation (BCS-1441502, CAREER PHY-1554488, BCS-1631550, and CNS-1626008). TDS was supported by R01MH107703. DRR was supported by K01MH102609. FP acknowledges support from NSF-BCS-1631112 and NSF-BCS-1430279. The content is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies. 17 References [1] Vicki A Anderson, Peter Anderson, Elisabeth Northam, Rani Jacobs, and Cathy Catroppa. Development of executive functions through late childhood and adolescence in an australian sample. Developmental neuropsychology, 20(1):385–406, 2001. [2] Daniel Romer, Laura Betancourt, Joan M. Giannetta, Nancy L. Brodsky, Martha Farah, and Hallam Hurt. Executive cognitive functions and impulsivity as correlates of risk taking and problem behavior in preadolescents. Neuropsychologia, 47(13):2916 – 2926, 2009. ISSN 0028-3932. doi: http://dx.doi.org/10.1016/j.neuropsychologia.2009.06.019. URL http://www. sciencedirect.com/science/article/pii/S002839320900270X. [3] Russell A Barkley, Kevin R Murphy, George J Dupaul, and Tracie Bush. Driving in young adults with attention deficit hyperactivity disorder: knowledge, performance, adverse out- comes, and the role of executive functioning. Journal of the International Neuropsychological Society, 8(5):655–672, 2002. [4] J. Biederman, C. R. Petty, R. Fried, A. E. Doyle, T. Spencer, L. J. Seidman, L. Gross, K. Poetzl, and S. V. Faraone. Stability of executive function deficits into young adult years: a prospective longitudinal follow-up study of grown up males with adhd. Acta Psychiatrica Scandinavica, 116(2):129–136, 2007. ISSN 1600-0447. doi: 10.1111/j.1600-0447.2007.01008.x. URL http://dx.doi.org/10.1111/j.1600-0447.2007.01008.x. [5] Constance L. Chapple and Katherine A. Johnson. Gender differences in impulsivity. Youth Violence and Juvenile Justice, 5(3):221–234, 2007. doi: 10.1177/1541204007301286. URL http://dx.doi.org/10.1177/1541204007301286. [6] Erik G Willcutt. The prevalence of dsm-iv attention-deficit/hyperactivity disorder: a meta- analytic review. Neurotherapeutics, 9(3):490–499, 2012. [7] Catharine P Cross, Lee T Copping, and Anne Campbell. Sex differences in impulsivity: a meta-analysis. Psychological bulletin, 137(1):97, 2011. [8] Sheik Hosenbocus and Raj Chahal. A review of executive function deficits and pharmacological management in children and adolescents. Journal of the Canadian Academy of Child and Adolescent Psychiatry, 21(3):223, 2012. [9] N Gogtay, J N Giedd, L Lusk, K M Hayashi, D Greenstein, A C Vaituzis, T F 3rd Nugent, D H Herman, L S Clasen, A W Toga, J L Rapoport, and P M Thompson. Dynamic mapping of human cortical development during childhood through early adulthood. Proc Natl Acad Sci U S A, 101(21):8174–8179, 2004. [10] G L Baum, R Ciric, D R Roalf, R F Betzel, T M Moore, R T Shinohara, A E Kahn, S N Vandekar, P E Rupert, M Quarmley, P A Cook, M A Elliott, K Ruparel, R E Gur, R C Gur, D S Bassett, and T D Satterthwaite. Modular segregation of structural brain networks supports the development of executive function in youth. Curr Biol, 27(11):1561–1572.e8, 2017. [11] V J Schmithorst, J Vannest, G Lee, L Hernandez-Garcia, E Plante, A Rajagopal, S K Holland, 18 and CMIND Authorship Consortium. Evidence that neurovascular coupling underlying the BOLD effect increases with age during childhood. Hum Brain Mapp, 36(1):1–15, 2015. [12] J S Nomi, T S Bolt, C E C Ezie, L Q Uddin, and A S Heller. Moment-to-moment BOLD signal variability reflects regional changes in neural flexibility across the lifespan. J Neurosci, 37(22):5539–5548, 2017. [13] E H Keulers, P Stiers, and J Jolles. Developmental changes between ages 13 and 21 years in the extent and magnitude of the BOLD response during decision making. Neuroimage, 54(2): 1442–1454, 2011. [14] D A Fair, A L Cohen, J D Power, N U Dosenbach, J A Church, F M Miezin, B L Schlaggar, and S E Petersen. Functional brain networks develop from a "local to distributed" organization. PLoS Comput Biol, 5(5):e1000381, 2009. [15] Efstathios D. Gennatas, Brian B. Avants, Daniel H. Wolf, Theodore D. Satterthwaite, Kosha Ruparel, Rastko Ciric, Hakon Hakonarson, Raquel E. Gur, and Ruben C. Gur. Age-related effects and sex differences in gray matter density, volume, mass, and cortical thickness from childhood to young adulthood. Journal of Neuroscience, 2017. ISSN 0270-6474. doi: 10. 1523/JNEUROSCI.3550-16.2017. URL http://www.jneurosci.org/content/early/2017/ 04/21/JNEUROSCI.3550-16.2017. [16] Sarah-Jayne Blakemore and Suparna Choudhury. Development of the adolescent brain: im- plications for executive function and social cognition. Journal of Child Psychology and Psy- chiatry, 47(3-4):296–312, 2006. ISSN 1469-7610. doi: 10.1111/j.1469-7610.2006.01611.x. URL http://dx.doi.org/10.1111/j.1469-7610.2006.01611.x. [17] Madhura Ingalhalikar, Alex Smith, Drew Parker, Theodore D. Satterthwaite, Mark A. Elliott, Kosha Ruparel, Hakon Hakonarson, Raquel E. Gur, Ruben C. Gur, and Ragini Verma. Sex differences in the structural connectome of the human brain. Proceedings of the National Academy of Sciences, 111(2):823–828, 2014. doi: 10.1073/pnas.1316909110. URL http:// www.pnas.org/content/111/2/823.abstract. [18] Theodore D Satterthwaite, Daniel H Wolf, David R Roalf, Kosha Ruparel, Guray Erus, Simon Vandekar, Efstathios D Gennatas, Mark A Elliott, Alex Smith, Hakon Hakonarson, et al. Linked sex differences in cognition and functional connectivity in youth. Cerebral cortex, 25 (9):2383–2394, 2014. [19] Y Y Liu, J J Slotine, and A L Barabasi. Controllability of complex networks. Nature, 473 (7346):167–173, 2011. [20] Fabio Pasqualetti, Sandro Zampieri, and Francesco Bullo. Controllability metrics, limitations and algorithms for complex networks. IEEE Transactions on Control of Network Systems, 1 (1):40–52, 2014. [21] T Kailath. Linear Systems. Prentice Hall, Englewood Cliffs, 1980. [22] R. E. Kalman, Y. C. Ho, and K. S. Narendra. Controllability of linear dynamical systems. Contributions to Differential Equations, 1:189–213, 1963. [23] S F Muldoon, F Pasqualetti, S Gu, M Cieslak, S T Grafton, J M Vettel, and D S Bassett. 19 Stimulation-based control of dynamic brain networks. PLoS Comput Biol, 12(9):e1005076, 2016. [24] R F Betzel, S Gu, J D Medaglia, F Pasqualetti, and D S Bassett. Optimally controlling the human connectome: the role of network topology. Sci Rep, 6:30770, 2016. [25] S Gu, R F Betzel, M G Mattar, M Cieslak, P R Delio, S T Grafton, F Pasqualetti, and D S Bassett. Optimal trajectories of brain state transitions. Neuroimage, 148:305–317, 2017. [26] L Wiles, S Gu, F Pasqualetti, D S Bassett, and D F Meaney. Autaptic connections shift network excitability and bursting. Scientific Reports, In Press, 2017. [27] J Kim, J M Soffer, A E Kahn, J M Vettel, F Pasqualetti, and D S Bassett. Role of graph architecture in controlling dynamical networks with applications to neural systems. Nature Physics, Epub Ahead of Print, 2017. [28] Shi Gu, Fabio Pasqualetti, Matthew Cieslak, Qawi K Telesford, B Yu Alfred, Ari E Kahn, John D Medaglia, Jean M Vettel, Michael B Miller, Scott T Grafton, et al. Controllability of structural brain networks. Nature communications, 6, 2015. [29] E Tang, C Giusti, G L Baum, S Gu, E Pollock, A E Kahn, D R Roalf, T M Moore, K Ruparel, R C Gur, R E Gur, T D Satterthwaite, and D S Bassett. Developmental increases in white matter network controllability support a growing diversity of brain dynamics. Nat Commun, 8(1):1252, 2017. [30] T D Satterthwaite, M A Elliott, K Ruparel, J Loughead, K Prabhakaran, M E Calkins, R Hopson, C Jackson, J Keefe, M Riley, F D Mentch, P Sleiman, R Verma, C Davatzikos, H Hakonarson, R C Gur, and R E Gur. Neuroimaging of the philadelphia neurodevelopmental cohort. Neuroimage, 86:544–553, 2014. [31] L Cammoun, X Gigandet, D Meskaldji, J P Thiran, O Sporns, K Q Do, P Maeder, R Meuli, and P Hagmann. Mapping the human connectome at multiple scales with diffusion spectrum MRI. J Neurosci Methods, 203(2):386–397, 2012. [32] David R Roalf, Megan Quarmley, Mark A Elliott, Theodore D Satterthwaite, Simon N Van- dekar, Kosha Ruparel, Efstathios D Gennatas, Monica E Calkins, Tyler M Moore, Ryan Hop- son, et al. The impact of quality assurance assessment on diffusion tensor imaging outcomes in a large-scale population-based cohort. Neuroimage, 125:903–919, 2016. [33] R C Gur, J Richard, P Hughett, M E Calkins, L Macy, W B Bilker, C Brensinger, and R E Gur. A cognitive neuroscience-based computerized battery for efficient measurement of individual differences: standardization and initial construct validation. J Neurosci Methods, 187(2):254–262, 2010. [34] Ruben C Gur, Jan Richard, Monica E Calkins, Rosetta Chiavacci, John A Hansen, Warren B Bilker, James Loughead, John J Connolly, Haijun Qiu, Frank D Mentch, et al. Age group and sex differences in performance on a computerized neurocognitive battery in children age 8- 21. Neuropsychology, 26(2):251, 2012. [35] Tyler M Moore, Steven P Reise, Raquel E Gur, Hakon Hakonarson, and Ruben C Gur. Psy- 20 chometric properties of the penn computerized neurocognitive battery. Neuropsychology, 29 (2):235, 2015. [36] Tyler M Moore, Steven P Reise, David R Roalf, Theodore D Satterthwaite, Christos Da- vatzikos, Warren B Bilker, Allison M Port, Chad T Jackson, Kosha Ruparel, Adam P Savitt, et al. Development of an itemwise efficiency scoring method: Concurrent, convergent, dis- criminant, and neuroimaging-based predictive validity assessed in a large community sample. Psychol Assess., 28(12):1529–1542, 2016. [37] Tyler M Moore, Ruben C Gur, Michael L Thomas, Gregory G Brown, Matthew K Nock, Adam P Savitt, John G Keilp, Steven Heeringa, Robert J Ursano, Murray B Stein, et al. Development, administration, and structural validity of a brief, computerized neurocognitive battery: Results from the army study to assess risk and resilience in service members. Assess- ment, Epub Ahead of Print, 2017. [38] Simon N Vandekar, Russell T Shinohara, Armin Raznahan, David R Roalf, Michelle Ross, Nicholas DeLeo, Kosha Ruparel, Ragini Verma, Daniel H Wolf, Ruben C Gur, et al. Topo- logically dissociable patterns of development of the human cerebral cortex. Journal of Neuro- science, 35(2):599–609, 2015. [39] B Fischl. Freesurfer. Neuroimage, 62:774–781, 2012. [40] Adon F.G. Rosen, David R. Roalf, Kosha Ruparel, Jason Blake, Kevin Seelaus, Lakshmi P. Villa, Rastko Ciric, Philip A. Cook, Christos Davatzikos, Mark A. Elliott, Angel Garcia de La Garza, Efstathios D. Gennatas, Megan Quarmley, J. Eric Schmitt, Russell T. Shi- nohara, M. Dylan Tisdall, R. Cameron Craddock, Raquel E. Gur, Ruben C. Gur, and Theodore D. Satterthwaite. Data-driven assessment of structural image quality. NeuroIm- age, pages –, 2017. ISSN 1053-8119. doi: https://doi.org/10.1016/j.neuroimage.2017.12.059. URL https://www.sciencedirect.com/science/article/pii/S1053811917310832. [41] Meritxell Bach Cuadra, Claudio Pollo, Anton Bardera, Olivier Cuisenaire, J-G Villemure, and J-P Thiran. Atlas-based segmentation of pathological mr brain images using a model of lesion growth. IEEE transactions on medical imaging, 23(10):1301–1314, 2004. [42] J Daniel Ragland, Bruce I Turetsky, Ruben C Gur, Faith Gunning-Dixon, Travis Turner, Lee Schroeder, Robin Chan, and Raquel E Gur. Working memory for complex figures: an fmri comparison of letter and fractal n-back tasks. Neuropsychology, 16(3):370, 2002. [43] Sheila Shanmugan, Daniel H Wolf, Monica E Calkins, Tyler M Moore, Kosha Ruparel, Ryan D Hopson, Simon N Vandekar, David R Roalf, Mark A Elliott, Chad Jackson, et al. Common and dissociable mechanisms of executive system dysfunction across psychiatric disorders in youth. American journal of psychiatry, 173(5):517–526, 2016. [44] D K Jones and P J Basser. "squashing peanuts and smashing pumpkins": how noise distorts diffusion-weighted MR data. Magn. Reson. Med., 52:979–993, 2004. [45] Y Chen, O Tymofiyeva, C P Hess, and D Xu. Effects of rejecting diffusion directions on tensor-derived parameters. Neuroimage, 109:160–170, 2015. [46] M Jenkinson, P Bannister, M Brady, and S Smith. Improved optimization for the robust and 21 accurate linear registration and motion correction of brain images. Neuroimage, 17:825–841, 2002. [47] J L R Andersson and S N Sotiropoulos. An integrated approach to correction for off-resonance effects and subject movement in diffusion MR imaging. Neuroimage, 125:1063–1078, 2016. [48] M Jenkinson, C F Beckmann, T E Behrens, M W Woolrich, and S M Smith. Fsl. Neuroimage, 62:782–790, 2012. [49] Theodore D Satterthwaite, Daniel H Wolf, Guray Erus, Kosha Ruparel, Mark A Elliott, Ef- stathios D Gennatas, Ryan Hopson, Chad Jackson, Karthik Prabhakaran, Warren B Bilker, et al. Functional maturation of the executive system during adolescence. Journal of Neuro- science, 33(41):16249–16261, 2013. [50] Douglas N Greve and Bruce Fischl. Accurate and robust brain image alignment using boundary-based registration. Neuroimage, 48(1):63–72, 2009. [51] CJ Honey, O Sporns, Leila Cammoun, Xavier Gigandet, Jean-Philippe Thiran, Reto Meuli, and Patric Hagmann. Predicting human resting-state functional connectivity from structural connectivity. Proceedings of the National Academy of Sciences, 106(6):2035–2040, 2009. [52] Louis M Pecora and Thomas L Carroll. Master stability functions for synchronized coupled systems. Physical review letters, 80(10):2109, 1998. [53] Yoav Benjamini and Yosef Hochberg. Controlling the false discovery rate: a practical and pow- erful approach to multiple testing. Journal of the royal statistical society. Series B (Method- ological), pages 289–300, 1995. [54] Simon N Wood. Stable and efficient multiple smoothing parameter estimation for generalized additive models. Journal of the American Statistical Association, 99(467):673–686, 2004. [55] Dustin Tingley, Teppei Yamamoto, Kentaro Hirose, Luke Keele, and Kosuke Imai. Mediation: R package for causal mediation analysis. 2014. [56] KJ Preacher, PJ Curran, and DJ Bauer. Simple intercepts, simple slopes, and regions of significance in mlr 2-way interactions. retrieved september 25, 2005, 2004. [57] Joan G Snodgrass and June Corwin. Pragmatics of measuring recognition memory: appli- cations to dementia and amnesia. Journal of Experimental Psychology: General, 117(1):34, 1988. [58] F Xavier Castellanos and Rosemary Tannock. Neuroscience of attention-deficit/hyperactivity disorder: the search for endophenotypes. Nature Reviews Neuroscience, 3(8):617–628, 2002. [59] Tiina Pohjalainen, Juha O Rinne, Kjell Nagren, Erkka Syvalahti, and Jarmo Hietala. Sex differences in the striatal dopamine d2 receptor binding characteristics in vivo. American Journal of Psychiatry, 155(6):768–773, 1998. [60] K.E. Stephan, F. Schlagenhauf, Q.J.M. Huys, S. Raman, E.A. Aponte, K.H. Brodersen, L. Rigoux, R.J. Moran, J. Daunizeau, R.J. Dolan, K.J. Friston, and A. Heinz. Compu- tational neuroimaging strategies for single patient predictions. NeuroImage, 145(Part B): ISSN 1053-8119. doi: https://doi.org/10.1016/j.neuroimage.2016.06.038. 180 – 199, 2017. 22 URL http://www.sciencedirect.com/science/article/pii/S1053811916302877. vidual Subject Prediction. Indi- [61] Eugene M Izhikevich and Gerald M Edelman. Large-scale model of mammalian thalamocortical systems. Proceedings of the national academy of sciences, 105(9):3593–3598, 2008. [62] A C Murphy and D S Bassett. A network neuroscience of neurofeedback for clinical translation. Curr Opin Biomed Eng, 1:63–70, 2017. [63] Jun Tanji and Eiji Hoshi. Role of the lateral prefrontal cortex in executive behavioral control. Physiological reviews, 88(1):37–57, 2008. [64] David R Euston, Aaron J Gruber, and Bruce L McNaughton. The role of medial prefrontal cortex in memory and decision making. Neuron, 76(6):1057–1070, 2012. [65] Cummings JL. Frontal-subcortical circuits and human behavior. Archives of Neurology, 50(8): 873–880, 1993. doi: 10.1001/archneur.1993.00540080076020. URL +http://dx.doi.org/10. 1001/archneur.1993.00540080076020. [66] Lisa M Duke and Alfred W Kaszniak. Executive control functions in degenerative dementias: A comparative review. Neuropsychology review, 10(2):75–99, 2000. [67] Raphael M Bonelli and Jeffrey L Cummings. Frontal-subcortical circuitry and behavior. Dia- logues in clinical neuroscience, 9(2):141, 2007. [68] Sibel Tekin and Jeffrey L Cummings. Frontal–subcortical neuronal circuits and clinical neu- ropsychiatry: an update. Journal of psychosomatic research, 53(2):647–654, 2002. [69] St´ephanie Bombois, St´ephanie Debette, Xavier Delbeuck, Am´elie Bruandet, Samuel Lepoit- tevin, Christine Delmaire, Didier Leys, and Florence Pasquier. Prevalence of subcortical vas- cular lesions and association with executive function in mild cognitive impairment subtypes. Stroke, 38(9):2595–2597, 2007. ISSN 0039-2499. doi: 10.1161/STROKEAHA.107.486407. URL http://stroke.ahajournals.org/content/38/9/2595. [70] Michael M. Plichta, Nenad Vasic, Robert Christian Wolf, Klaus-Peter Lesch, Dagmar Brum- mer, Christian Jacob, Andreas J. Fallgatter, and Georg Gron. Neural hyporesponsiveness and hyperresponsiveness during immediate and delayed reward processing in adult attention- deficit/hyperactivity disorder. Biological Psychiatry, 65(1):7 – 14, 2009. ISSN 0006-3223. doi: https://doi.org/10.1016/j.biopsych.2008.07.008. URL http://www.sciencedirect.com/ science/article/pii/S0006322308008275. Perception, Empathy, and Reward in Attention- Deficit/Hyperactivity Disorder and Autism. [71] Ana Cubillo, Rozmin Halari, Anna Smith, Eric Taylor, and Katya Rubia. A review of fronto- striatal and fronto-cortical brain abnormalities in children and adults with attention deficit hyperactivity disorder (adhd) and new evidence for dysfunction in adults with adhd during motivation and attention. Cortex, 48(2):194 – 215, 2012. ISSN 0010-9452. doi: https://doi. org/10.1016/j.cortex.2011.04.007. URL http://www.sciencedirect.com/science/article/ pii/S001094521100102X. Frontal lobes. [72] Chandan J Vaidya. Neurodevelopmental abnormalities in adhd. In Behavioral neuroscience of attention deficit hyperactivity disorder and its treatment, pages 49–66. Springer, 2011. 23 [73] Naftali Raz, Ivan J Torres, and James D Acker. Age, gender, and hemispheric differences in human striatum: a quantitative review and new data from in vivo mri morphometry. Neuro- biology of learning and memory, 63(2):133–142, 1995. [74] Walter J. Freeman and Christine A. Skarda. Spatial eeg patterns, non-linear dynamics the neo-sherringtonian view. Brain Research Reviews, 10(3):147 – 175, ISSN 0165-0173. doi: https://doi.org/10.1016/0165-0173(85)90022-0. URL http: and perception: 1985. //www.sciencedirect.com/science/article/pii/0165017385900220. [75] Tong Zhu, Rui Hu, Xing Qiu, Michael Taylor, Yuen Tso, Constantin Yiannoutsos, Bradford Navia, Susumu Mori, Sven Ekholm, Giovanni Schifitto, et al. Quantification of accuracy and precision of multi-center dti measurements: a diffusion phantom and human brain study. Neuroimage, 56(3):1398–1411, 2011. [76] Klaus H Maier-Hein, Peter F Neher, Jean-Christophe Houde, Marc-Alexandre Cot´e, Eleftherios Garyfallidis, Jidan Zhong, Maxime Chamberland, Fang-Cheng Yeh, Ying-Chia Lin, Qing Ji, et al. The challenge of mapping the human connectome based on diffusion tractography. Nature communications, 8:1349, 2017. [77] Marek Kubicki, Robert McCarley, Carl-Fredrik Westin, Hae-Jeong Park, Stephan Maier, Ron Kikinis, Ferenc A Jolesz, and Martha E Shenton. A review of diffusion tensor imaging studies in schizophrenia. Journal of psychiatric research, 41(1):15–30, 2007. [78] Michela Pievani, Willem de Haan, Tao Wu, William W Seeley, and Giovanni B Frisoni. Func- tional network disruption in the degenerative dementias. The Lancet Neurology, 10(9):829–843, 2011. [79] Daphna Joel, Zohar Berman, Ido Tavor, Nadav Wexler, Olga Gaber, Yaniv Stein, Nisan Shefi, Jared Pool, Sebastian Urchs, Daniel S Margulies, et al. Sex beyond the genitalia: The human brain mosaic. Proceedings of the National Academy of Sciences, 112(50):15468–15473, 2015. [80] Daphna Joel and Margaret M McCarthy. Incorporating sex as a biological variable in neu- ropsychiatric research: where are we now and where should we be? Neuropsychopharmacology, 42(2):379–385, 2017. 24 Figures Figure 1: Control theory and schematic of data processing. (a) Schematic depicting ba- sic methodological approach. Diffusion tensor imaging (DTI) data was acquired from 882 youth ages 8 to 22 years. Deterministic tractography was used to identify the number of white matter streamlines, connecting any two regions of interest. These estimates were used to construct a struc- tural brain network for each subject representing white matter connectivity (edges) between brain regions (nodes). (b) Diagram illustrating brain state transitions from an activated default mode system (Initial ). The blue arrow denotes a distant transition to a deactivated default mode system and activated frontoparietal/dorsal attention system (Distant). The red arrow denotes a nearby transition to a partially deactivated default mode system (Near ). Regions with high modal con- trollability can facilitate transitions to energetically distant states while regions with high average controllability can facilitate transitions to nearby states while requiring very little energy. 25 a.DTIDeterministicTractographyConnectivity MatrixBrain Network0Energy1b.InitialDistantNearModalcontrollabilityAveragecontrollability Figure 2: Cognitive development and network controllability by sex. (a) Executive effi- ciency, a summary of performance on tasks taxing executive function in the PCNB, fit non-linearly on age with a penalized regression spline. Red and blue data points and curves represent females and males, respectively. Each point represents a subject's partial residual with respect to sex; solid lines represent the predicted mean for males and females separately, and shaded envelopes denote 95% confidence intervals of the prediction. (b) Rank of mean average controllability values at each node across 389 male subjects and 493 female subjects, separately. The Spearman rank correlation between regional average controllability averaged across males and regional average controllability averaged across females was r = 0.9915. 26 a.b.r = 0.99MalesFemales1234−3−2−1012810121416182022Age (y)Executive Efficiency Figure 3: Sex differences in regional controllability. (a-b) Controllability estimates averaged over 219 cortical nodes (a) and 14 subcortical nodes (b), for average controllability and modal controllability. Females have higher cortical average controllability, cortical modal controllabillity and subcortical modal controllability, but males have higher subcortical average controllability. In panels a-b, red and blue represent females and males, respectively. Each point represents a subject's partial residual with respect to sex; bar height represents the predicted mean for males and females separately, and shaded grey envelopes denote 95% confidence intervals of the prediction. (c-d) Heatmaps where color intensity corresponds to standardized regression coefficient for sex, with warmer colors indicating higher controllability in females. Average controllability is depicted in (c) and modal controllability is depicted in (d). Colors reflect values of standardized regression coefficients for controllability at each node as a predictor of executive function while controlling for covariates (Eq. 5). *, p < 0.05, **, p < 0.01, ***, p < 10−4. 27 c.a.b.-0.2100.15M> FM > F0F > MF > M101520FemalesMalesCorticalAverage Controllability0.940.960.981.00FemalesMalesCorticalModal Controllability100200FemalesMalesSubcorticalAverage Controllability0.900.951.00FemalesMalesSubcorticalModal Controllabilityd.Average Controllability ~ SexModal Controllability ~ Sex0.18-0.18zz Figure 4: Network controllability as a function of age. (a-d) Relationship between age and controllability estimates, averaged over 219 cortical nodes (a-b) and 14 subcortical nodes (c-d), for average controllability (a, c) and modal controllability (b, d). Modal controllability increases with age in the cortex and in the subcortex; average controllability increases with age in the cortex. In the subcortex, there was a significant age-by-sex interaction for average controllability. However, the simple slopes for males (βage) and females (βage + βage×sex) were not significantly different from 0. Age-by-sex interactions in a, b, and d were not significant. Red and blue represent females and males, respectively. Each point represents a subject's partial residual with respect to age; lines represent regression slopes for males and females separately, and shaded envelopes denote 95% confidence intervals of the slopes. The p values for βage are shown in panels (a,b,d) and for βage×sex are shown in panel (c). 28 a.b.c.d.FemalesMales8121620pβage<10e−41015208121620Age (y)Average ControllabilityCorticalSubcorticalpβint<0.051002008121620Age (y)Average Controllabilitypβage<10e−40.900.951.008121620Age (y)Modal Controllabilitypβage<10e−40.940.960.981.00Age (y)Modal Controllability Figure 5: Sex effects on relationships between metrics of network control and dynamics. Relationships between whole-brain average and modal controllability (a, e), between synchroniz- ability and whole-brain average controllability (d, h), and between regional average controllability and regional modal controllability (b-c, f-g). Data from male subjects is depicted in panels a-d and data from female subjects is depicted in panels e-h. Non-linear least squares was used on the base model: y = a+b∗exp(c∗x), and the full model: y = (a+α∗Sex)+(b+β∗Sex)∗exp((c+γ∗Sex)∗x). Slopes represent coefficients from the full model. The p values were obtained from ANOVA com- paring base and full models. In these cases, a low p value indicates that allowing for a sex term in the model significantly increases the explained variance. *, p < 0.05, **, p < 0.01, ***, p < 10−4. 29 912151821Age (y)a.MalesFemalesb.c.d.h.g.e.0.940.950.960.970.98202530Modal ControllabilityGlobal Mean***************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************0.950.960.970.98101520Modal ControllabilityCortical Mean***************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************0.8750.9000.9250.9500.975100200Modal ControllabilitySubcortical Mean1.01.52.02.5202530SynchronizabilitySynchronizability0.940.950.960.970.98202530Average ControllabilityModal Controllability***************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************0.950.960.970.98101520Average ControllabilityModal Controllability***************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************************0.8750.9000.9250.9500.975100200Average ControllabilityModal Controllability1.01.52.02.5202530Average ControllabilitySynchronizabilityf. Figure 6: Network controllability and cognitive performance. (a-d) Relationship between age and controllability estimates, averaged over 219 cortical nodes (a-b) and 14 subcortical nodes (c-d), for average controllability (a, c) and modal controllability (b, d). Cortical controllability (a-b) is not significantly related to executive function, but subcortical average controllability (c) is negatively correlated with executive function while subcortical modal controllability (d) is posi- tively correlated with executive function. In panels a-d, red and blue represent females and males, respectively. Each data point represents a subject's partial residual with respect to controllability; lines represent regression slopes for males and females separately, and shaded envelopes denote 95% confidence intervals of the slopes. (e) Brain regions with sex-associated average controllability and their relation to executive function (FDR corrected, q < 0.05). Colors reflect values of standardized regression coefficients for average controllability at each node as a predictor of executive function while controlling for covariates. Modal controllability was not significantly associated with execu- tive function. For panels a-e, there were no significant sex-controllability interactions on executive function. 30 a.b.c.d.e.-0.080.080zExecutive Efficiency ~ Average ControllabilityNS−3−2−1012101520Average ControllabilityExecutive EfficiencyNS−3−2−10120.960.981.00Modal Controllabilitypβage<0.01−3−2−101250100150200250Average ControllabilityExecutive Efficiencypβage<0.05−3−2−10120.880.920.961.00Modal ControllabilityCorticalSubcortical R Medial orbito frontal, parcel 1 R Superior frontal, parcel 4 R Postcentral, parcel 5 R Caudate L Caudate Average Controllability Control Node ACME 0.03 0.01 0.01 0.03 0.02 p ADE 0.13 0.15 0.15 0.13 0.14 0.00 0.43 0.10 0.01 0.03 p 0.01 0.01 0.00 0.01 0.01 Figure 7: Average controllability mediates the relation between sex and executive func- tion. (a) Non-parametric causal mediation analysis for sex → average controllability → executive function, using control nodes with sex-associated average controllability values. ACME = average causal mediation effect, i.e. average controllability → executive function; ADE = average direct effect, i.e. sex → executive function. Associations between sex and controllability from Fig. 3c are shown; cooler colors indicate higher controllability in males. 1000 simulations were performed. p-values were FDR corrected (q < 0.05). Age, total brain volume, handedness, and head motion were included in all analyses. * indicates standardized regression coefficient for sex as a predictor of executive function with the same covariates in the model. 31 SexExecutive Efficiency(Fig. 2c)-0.030.030ControllabilityAverage-0.210.150ADE0.16* Figure 8: Regional controllability predicts n-back task activation and cognitive per- formance differently for males and females. (a) Brain regions at which the 2-back minus 0-back GLM β weights are associated with 2-back d(cid:48), a summary measure of task performance in n = 659 subjects with quality fMRI data. (b) Heatmap depicting standardized multiple linear regression β's for regional average controllability as a predictor of regional 2-back minus 0-back activation. Only nodes with associations surviving FDR correction are shown (q < 0.05). (c) Vi- sual representation of the data presented in panel (b); average controllability at highlighted nodes is associated with activation at nodes of the same colors. (d) Associations between activation and average controllability for n = 283 males and n = 376 females, separately. Modal controllability was not significantly associated with activation. For panels (c,d), corresponding colors indicate association between control node and regional activation. 32 -0.200.2a.b.c.d.d' ~ Activation MalesFemales-0.1600.22zzControllabilityActivation A Supplementary Data 33 Figure A.1: Whole-brain network controllability and synchronizability by sex. (a, c) Average (a) and modal (c) controllability, averaged across all brain regions (hereafter "global"), comparing males and females. (b) Whole-brain ("global") synchronizability over age. (d) Global synchronizability by sex. We observed no significant age-by-sex interactions on synchronizability. Red and blue represent females and males, respectively. Each point represents a subject's partial residual with respect to sex; bar height represents the predicted mean for males and females sep- arately, and shaded envelopes denote 95% confidence intervals of the prediction. *, p < 0.05, **, p < 0.01, ***, p < 10−4. 34 **1520253035a.b.c.d.FemalesMalesGlobal Mean Average Controllabilitypβage<10e−161.01.52.02.5810121416182022Age (y)Global Synchronizability**0.920.940.960.981.00FemalesMalesGlobal Mean Modal ControllabilityNS1.01.52.02.5FemalesMalesGlobal SynchronizabilityFemalesMales Figure A.2: Trends between regional weighted degree, age, and sex differ from those with network controllability. (a-d) Weighted degree averaged over 219 cortical nodes (a-b) and 14 subcortical nodes (c-d). Panels (a, c) show that no significant sex differences exist in mean cortical (a) or subcortical (c) weighted degree. Panels (b, d) show that weighted degree decreases with age in the cortex (b) and increases with age in the subcortex (d). We observed no significant age-by-sex interactions in these models. Red and blue represent females and males, respectively. Each point represents a subject's partial residual with respect to sex; bar height represents the predicted mean for males and females separately, and shaded envelopes denote 95% confidence intervals of the prediction. *, p < 0.05, **, p < 0.01, ***, p < 10−4. 35 NS3500400045005000FemalesMalesMean Cortical Weighted Degreepβage<0.013500400045005000810121416182022Age (y)Mean Cortical Weighted DegreeNS5000700090001100013000FemalesMalesMean Subcortical Weighted Degreepβage<0.055000700090001100013000810121416182022Age (y)Mean Subcortical Weighted Degreea.b.c.d.FemalesMales Figure A.3: Relationship between sex, network controllability, and cognitive perfor- mance is not well-explained by weighted degree. (a) Heatmaps where color intensity corre- sponds to standardized regression coefficient for sex, with warmer colors indicating higher weighted degree in females. Five out of twenty-one regions with sex-associated controllability also have sex- associated weighted degree. Of those five, only average controllability values at the left caudate nucleus and right postcentral gyrus have the same direction of assocation with sex as weighted de- gree. (b) Association between executive function and weighted degree at nodes with sex-associated average controllability, modal controllability, or weighted degree, correcting for multiple compar- isons with an FDR correction at q < 0.05. No sex × weighted degree interaction terms on executive function were significant. Age, total brain volume, handedness, and head motion were included as covariates in all analyses. 36 b.-0.0700.07Executive Efficiency ~ Weighted Degreea.-0.150 Figure A.4: Sex-dependent relationships between network controllability and activation. (a) Heatmap of standardized multiple linear regression β's for regional average controllability as a predictor of regional 2-back minus 0-back activation. (Left) Data from n = 376 females. (Right) Data from n = 283 males. Only nodes with associations surviving FDR correction are shown (q < 0.05). (b) Heatmap of standardized β's for average controllability × sex interaction. No β was significant after FDR correction (q < 0.05), but nodes from panel (a) are shown to illustrate consistent directionality of association. Especially obvious are the left caudal middle frontal (parcel 1) and left banks of STS (parcel 2), where darker colors on the interaction heatmap indicate a larger effect size in males. The left fusiform gyrus (parcel 4) shows the opposite trend, with lighter colors on the interaction heatmap indicating a larger effect size in females. 37 a.b.-0.50.5-0.20.200-0.250.250F > MM > Fzz Figure A.5: Relationship between weighted degree and activation. (a) Heatmap of multiple linear regression β's for regional weighted degree as a predictor of regional 2-back minus 0-back activation. Only nodes with associations surviving FDR correction are shown (q < 0.05). There were only significant associations in the sample of n = 283 males, and not in the pooled sample or in females alone. (b) Visualization of associations between weighted degree and activation. Weighted degree at highlighted nodes is associated with activation at nodes of the same colors; that is, corresponding colors indicate the association between structural connections and regional activation. 38 Weighted DegreeActivationMalesb.0.150-0.15za. R Medial orbito frontal1 R Rostral middle frontal6 R Superior frontal4 R Postcentral5 R Superior parietal1 R Inferior parietal1 R Inferior parietal3 R Lateral occipital1 R Fusiform4 R Caudate L Lateral orbitofrontal3 L Pars triangularis1 L Precentral4 L Precentral8 L Postcentral1 L Postcentral3 L Postcentral4 L Superior parietal2 L Superior parietal5 L Inferior parietal3 L Lateral occipital4 L Lateral occipital5 L Fusiform4 L Superior temporal4 L Insula3 L Caudate L Putamen L Pallidum L Amygdala ACME 0.01 0.01 0.00 0.01 -0.00 -0.00 -0.00 -0.01 0.00 0.01 -0.01 -0.00 0.00 0.00 -0.00 0.00 -0.00 -0.01 -0.00 0.00 0.00 0.00 0.01 -0.01 0.00 0.01 0.00 0.01 0.01 a. Weighted Degree p ADE 0.15 0.15 0.15 0.15 0.16 0.15 0.16 0.16 0.15 0.15 0.17 0.16 0.16 0.16 0.16 0.15 0.16 0.17 0.16 0.16 0.16 0.15 0.15 0.16 0.15 0.15 0.15 0.15 0.15 0.54 0.25 0.79 0.77 0.91 0.91 0.91 0.54 0.85 0.06 0.25 0.91 0.91 0.91 0.91 0.91 0.91 0.54 0.91 0.86 0.91 0.86 0.78 0.70 0.86 0.54 0.70 0.61 0.54 p Total Effect 0.16 0.16 0.15 0.15 0.15 0.15 0.15 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.15 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.16 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 p Prop. Mediated 0.06 0.06 0.02 0.05 -0.00 -0.00 -0.01 -0.04 0.03 0.08 -0.07 -0.00 0.01 0.01 -0.01 0.00 -0.01 -0.03 -0.01 0.01 0.00 0.02 0.05 -0.04 0.01 0.06 0.02 0.04 0.04 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 p 0.55 0.27 0.79 0.77 0.91 0.91 0.91 0.55 0.86 0.12 0.27 0.91 0.91 0.91 0.91 0.91 0.91 0.55 0.91 0.86 0.91 0.86 0.78 0.70 0.86 0.55 0.70 0.61 0.55 Figure A.6: Sex-(weighted degree)-executive function mediation relationship. (a-b) Non- parametric causal mediation analysis for sex → weighted degree → executive function, using control nodes with sex-associated average controllability, modal controllability, or weighted degree. ACME = average causal mediation effect, ADE = average direct effect. We performed 1000 simulations. The p-values were FDR corrected (q < 0.05). Age, total brain volume, handedness, and head motion were included in all analyses. 39
1112.4987
1
1112
2011-12-21T11:07:16
Biologically-Inspired Electronics with Memory Circuit Elements
[ "q-bio.NC", "cond-mat.mes-hall" ]
Several abilities of biological systems, such as adaptation to natural environment, or of animals to learn patterns when appropriately trained, are features that are extremely useful, if emulated by electronic circuits, in applications ranging from robotics to solution of complex optimization problems, traffic control, etc. In this chapter, we discuss several examples of biologically-inspired circuits that take advantage of memory circuit elements, namely, electronic elements whose resistive, capacitive or inductive characteristics depend on their past dynamics. We provide several illustrations of what can be accomplished with these elements including learning circuits and related adaptive filters, neuromorphic and cellular computing circuits, analog massively-parallel computation architectures, etc. We also give examples of experimental realizations of memory circuit elements and discuss opportunities and challenges in this new field.
q-bio.NC
q-bio
Biologically-Inspired Electronics with Memory Circuit Elements Massimiliano Di Ventra and Yuriy V. Pershin Abstract Several unique properties of biological systems, such as adaptation to nat- ural environment, or of animals to learn patterns when appropriately trained, are features that are extremely useful, if emulated by electronic circuits, in applications ranging from robotics to solution of complex optimization problems, traffic con- trol, etc. In this chapter, we discuss several examples of biologically-inspired cir- cuits that take advantage of memory circuit elements, namely, electronic elements whose resistive, capacitive or inductive characteristics depend on their past dynam- ics. We provide several illustrations of what can be accomplished with these ele- ments including learning circuits and related adaptive filters, neuromorphic and cel- lular computing circuits, analog massively-parallel computation architectures, etc. We also give examples of experimental realizations of memory circuit elements and discuss opportunities and challenges in this new field. 1 Introduction Reproducing some of the features that are commonly found in living organisms, in- cluding - as the ultimate, and most sought after goal - the workings of the human brain, is what "artificial intelligence" is all about [1]. However, even without aiming for such an ambitious target, there are several tasks that living organisms perform seamlessly and, when reproduced in electronic circuits, are of great benefit and help us solve complicated problems. The main idea of biologically-inspired electronics is Massimiliano Di Ventra Department of Physics, University of California San Diego, La Jolla, CA 92093 USA, e-mail: [email protected] Yuriy V. Pershin Department of Physics and Astronomy and USC Nanocenter, University of South Carolina, Columbia, SC 29208 USA, e-mail: [email protected] 1 2 Massimiliano Di Ventra and Yuriy V. Pershin thus in borrowing approaches used by biological systems interacting with their envi- ronment and in applying them in diverse technological areas requiring, for example, adaptation to changing inputs, analog solution of optimization problems -- one of the well-known approaches to this problem is the "ant-search algorithm" [2, 3] -- , associative memory and unlearning, binary and fuzzy logic, etc. If we look closer, it is evident that if we want to reproduce any of these tasks with electronic circuits, two main characteristics have to be satisfied by some of the circuit elements or their combinations. These elements need to i) store information -- they have to have memory of the past -- and ii) be dynamical -- their states have to vary in time in response to a given input, preferably in a non-linear way. The latter requirement will help building a wide range of electronic circuits of desired functionality. Circuits based on active elements (such as transistors) can clearly perform both tasks, however, at a high cost of power consumption, low density, and complexity. It would be much more desirable if we could combine the above features in single, passive elements, preferably with dimensions at the nanometer scale, and hence comparable to -- or even smaller than -- biological storing and processing units, such as synapses and neurons. Such elements do exist, and go under the name of memristive, memcapacitive and meminductive systems, or collectively simply named memelements [4]. These are resistors, capacitors and inductors, respectively, whose state at a given time de- pends on the history of applied inputs (e.g., charge, voltage, current, or flux) and states through which the system has evolved. As we have shown, these memele- ments provide an unifying description of materials and systems with memory [5], in the sense that all two-terminal electronic devices based on memory materials and systems, when subject to time-dependent perturbations, behave simply as -- or as a combination of -- memristive, memcapacitive and meminductive systems, namely, dynamical non-linear circuit elements with memory [6]. In this Chapter, we will show how the analog memory features of memelements are ideal to reproduce a host of processes typical of living organisms. We will review mainly work by the authors in various contexts, ranging from learning (adaptive) cir- cuits to associative memory, and inherent massive parallelism afforded by networks of memelements. The Chapter is then organized as follows. Section 2 briefly reviews both the definition of these memory circuit elements and their main properties (a full account can be found in the original publications [7, 8, 4] and in our recent review paper [6]). Several experimental realizations exemplifying memristive, memcapac- itive and meminductive systems are discussed in Sec. 3. Section 4 is devoted to biologically-inspired circuits based on memory circuit elements including simple adaptive circuits (Sec. 4.1), neuromorphic circuits (Sec. 4.2) and massively-parallel analog processing circuits (Sec. 4.3). Concluding remarks are given in Sec. 5. Biologically-Inspired Electronics with Memory Circuit Elements 3 Fig. 1 Symbols of memory circuit elements: memristor, memcapacitor and meminductor. Gener- ally, memelements are asymmetric devices. The thick horizontal lines above the bottom electrodes are employed to define the device polarity [4]. Reprinted with permission from Ref. [4]. c(cid:13)2009 IEEE. 2 Definitions and properties Let us consider electronic devices defined by all possible pairs of fundamental cir- cuit variables u(t) and y(t) (i.e., current, charge, voltage, or flux). For each pair, we can introduce a response function, g, that, generally, depends on a set of n state variables, x = {xi}, i = 1, ..,n, describing the internal state of the system [4] 1. For instance, resistance of certain systems may depend on spin polarization [9, 10] or temperature [8]; capacitance and inductance of some other elements may exhibit dependence on system geometry [11, 6] or electric polarization [12]. In general, all internal microscopic physical properties related to the memory response of these electronic devices should be included into the vector of internal state variables x. The resulting memory circuit elements are then described by the following relations [4] y(t) = g (x,u,t)u(t) x = f (x,u,t) (1) (2) where f is a continuous n-dimensional vector function. Eqs. (1), (2) have to be sup- plied by appropriate initial conditions [13]. If u is the current and y is the voltage then Eqs. (1), (2) define memristive (for memory resistive) systems. In this case g is the memristance (memory resistance). In memcapacitive (memory capacitive) sys- tems, the charge q is related to the voltage V so that g is the memcapacitance (mem- ory capacitance). Finally, in meminductive (memory inductive) systems the flux ϕ is related to the current I with g the meminductance (memory inductance). There are still three additional pairs of fundamental circuit variables. However, these do not give rise to any new devices. For example, the pairs charge-current and voltage- flux are linked through equations of electrodynamics. Moreover, we could redefine devices defined by the charge-flux (which is the integral of the voltage) pair in the current-voltage basis [7]. Circuits symbols of memristive, memcapacitive and me- minductive systems are presented in Fig. 1. 1 There is no such dependence for traditional basic circuit elements -- resistors, capacitors and inductors. MemristorMemcapacitorMeminductorMemristorMemcapacitorMeminductor 4 Massimiliano Di Ventra and Yuriy V. Pershin Let us consider memristive systems in more details (definitions of all the other elements can be easily derived by considering the different constitutive variables [4]). Specifically, a current-controlled memristive system [8, 4] is defined by Eqs. (1), (2) as VM(t) = R (x,I,t)I(t), x = f (x,I,t) , (3) (4) where VM(t) and I(t) denote the voltage and current across the device, and R is the memristance. In a simple model of an ideal memristor [7], the memristance depends only on charge -- the time integral of the current. One particular realization of such a model has been suggested in Ref. [14] and is formulated as i j = RONx + ROFF (1− x) , RM (5) where RON and ROFF are minimal and maximal values of memristance, and x is a dimensionless internal state variable bound to the region 0 ≤ x ≤ 1. The dynamics of x can then be simply chosen as [14] dx dt = αI(t), (6) where α is a constant and I(t) is the current flowing through the memristor. Another example of memristive systems (which we will make use of later in this chapter) is a threshold-type memristive system [15]. Its model is specified by a threshold voltage (and some other parameters) that defines different device response regions (with regard to the voltage applied across the device). Mathematically, the threshold-type memristive system is described by the following equations I = x−1VM, x = (βVM + 0.5 (α − β ) [VM +Vt−VM −Vt]) ×θ (x/R1 − 1)θ (1− x/R2) , (7) (8) where I and VM are the current through and the voltage drop on the device, respec- tively, and x is the internal state variable playing the role of memristance, R = x, θ (·) is the step function, α and β characterize the rate of memristance change at VM ≤ Vt and VM > Vt, respectively, Vt is a threshold voltage, and R1 and R2 are limiting values of the memristance R. In Eq. (8), the θ-functions symbolically show that the memristance can change only between R1 and R2. On a practical level, the value of x must be monitored at each time step and in the situations when x < R1 or x > R2, it must be set equal to R1 or R2, respectively. In this way, we avoid situations when x may overshoot the limiting values by some amount and thus not change any longer because of the step function in Eq. (8). We have introduced and employed this model to describe the learning properties of unicellular organisms, associative memory, etc., as we will describe in some detail in the following sections. Biologically-Inspired Electronics with Memory Circuit Elements 5 There are several properties that characterize memory circuit elements. We refer the reader to the original papers [7, 8, 4] and the extensive review [6] where these are discussed at length. Here, we just mention that they are typically characterized by a frequency-dependent "pinched hysteresis loop" in their constitutive variables when subject to a periodic input. Also, normally, the memristance, memcapacitance and meminductance acquire values between two limits (with exceptions as discussed in Refs. [6, 12, 16]). Although the hysteresis of these elements under a periodic input may strongly depend on initial conditions [13], it is generally more pronounced at frequencies of the external input that are comparable to frequencies of internal processes that lead to memory. In many cases, at very low frequencies memory circuit elements behave as non-linear elements while at high frequencies as linear elements. Also, the hysteresis loops may or may not show self-crossing - which we have named type-I and type-II crossing behavior, respectively [6] - and often the internal state variable remains unchanged for a long time without any input signal applied. This provides non-volatile memory, which is an important feature for some of the applications we discuss later. Finally, we mention that the state variables - whether from a continuum or a discrete set of states - may follow a stochastic differential equation rather than a de- terministic one [6]. Interesting effects have been predicted in the presence of noise, such as noise-induced hysteresis [17]. This may have a large bearing in simulating biological processes - which necessarily occur under noisy conditions - and could be used to enhance the performance of certain devices. 3 Experimental realizations In this section, we briefly discuss experimental realizations of memory circuit el- ements. There is a large amount of experimental systems showing memristive be- havior (based, however, on very different physical mechanisms). For example, in thermistors -- being among the first identified memristive systems [8] -- the memory effects are related to thermal effects, mainly on how fast the heat dissipation occurs. In spintronics memristive systems, either based on semiconductor [9] or metal [10] materials, the memory feature is provided by the electron spin degree of freedom. Finally, resistance switching memory cells are probably the most important type of memristive systems. These cells are normally built of two metal electrodes sepa- rated by a gap of a few tens of nanometers filled by a memristive material. Different physical processes inside the memristive material can be responsible for the mem- ory. Fig. 2 shows an example of resistance switching memory cell -- an electrochem- ical metallization cell -- in which a layer of dielectric material (SiO2) separates two dissimilar metal electrodes (made of copper and platinum) [18]. Externally applied voltage induces migration of copper atoms that can bridge the gap between the elec- trodes thus reducing the cell's resistance. Such a bridge can also be disrupted by the applied voltage of the opposite polarity (see Fig. 2). 6 Massimiliano Di Ventra and Yuriy V. Pershin Fig. 2 Current-voltage char- acteristic of a Cu/SiO2/Pt electrochemical metallization cell recorded using a triangu- lar voltage sweep. The insets show dynamics of metallic fil- ament formation. (Reprinted with permission from [18]. 2009 American Institute of Physics.) Memcapacitive effects can be related to changes in the capacitor geometry or specific intrinsic properties of dielectric medium located between the capacitor plates [6]. For instance, the former mechanism plays the main role in elastic mem- capacitive systems that could be based on internal elastic states of the capacitor plate (e.g., direction of bending [11]) or elasticity of a medium between the plates that could be thought of as a spring [6]. Examples of the latter mechanism include memcapacitive systems with a delayed polarization response [12, 16] and struc- tures with permittivity switching [19]. Although meminductive systems are the least studied memory circuit elements at the moment, there are several known systems showing such type of functionality. For instance, in bimorph meminductive sys- tems [20, 21, 22], the inductance depends on the inductor's shape defined by the inductor temperature. Heat dissipation mechanisms play a significant role in this type of systems. Many additional examples of memristive, memcapacitive and me- minductive systems can be found in our recent review paper [6]. The most important aspect of all these examples is that they relate primarily to structures with nanoscale dimensions. This is not surprising since (up to a certain limit) the smaller the di- mensions of the system the easier it is to observe memory effects. In addition, we would like to mention that several types of memory effects may be present in a single device. For example, the coexistance of memristive and mem- capacitive responses has been observed in perovskite oxide thin films [23]. More- over, several three-terminal transistor-like memristive devices have been investi- gated in the past [24, 25, 26]. Clearly, different types of memory materials can be combined to obtain multi-terminal device structures with complex functionalities. Biologically-Inspired Electronics with Memory Circuit Elements 7 Fig. 3 (a) A possible real- ization of the learning circuit consisting of resistor R, in- ductor L, capacitor C and memristive system M. (b) Schematics of the function f (V ) defining a threshold- type memristive system (see also Eqs. (7), (8)). Reprinted with permission from Ref. [15]. c(cid:13)2009 American Phys- ical Society. 4 Biologically-Inspired Circuits 4.1 Modeling adaptive behavior of unicellular organisms Adaptive behavior is common to all life forms on Earth of all five kingdoms of na- ture: in Plantae (the plants) [27, 28, 29], Animalia (the animals) [30], Protista (the single-celled eukaryotes) [31, 32], Fungi (fungus and related organisms) [33, 34], and Monera (the prokaryotes) [35, 36, 37]. (The literature on this subject is exten- sive, hence we have given only a few representative references.) To a greater or lesser extent, representatives of all life forms respond (adapt) to changes of their en- vironment in a manner that increases the survival of their species. It would thus be of benefit to mimic and use this important natural feature in artificial structures, in particular in electronics to allow novel functionalities otherwise nonexistent in stan- dard circuitry. There are indeed many domains where such learning circuits can be employed. These range from robot control systems to signal processing. Therefore, developing circuit models of the adaptive behavior of the natural world is of great importance in many scientific areas [38]. Note that by "learning" here we simply mean the ability to adapt to incoming signals with retention of such information for later use. In this section we consider a particularly interesting example: the ability of the slime mold Physarum polycephalum to adapt its locomotion to periodic changes of the environment [32]. The simplicity of the system - a unicellular organism - and its well-defined response to specific input signals, make it an ideal test bed for the application of the notion of memory circuit elements in biology, and a source of in- spiration for more complex adaptive behavior in living organisms. In addition, this is a particularly appealing example of the full range of properties of memelements, in particular, their analog capability, which expands their range far beyond the digital domain. V(t)Rf(b)(a)LVVTVCM(t)V-VTC 8 Massimiliano Di Ventra and Yuriy V. Pershin In particular, it has been shown in a recent experiment [32] that the Physarum polycephalum subjected to periodic unfavorable environmental conditions (lower temperature and humidity) not only responds to these conditions by reducing its locomotion speed, but also anticipates future unfavorable environmental conditions reducing the speed at the time when the next unfavorable episode would have oc- curred. While the microscopic mechanism of such behavior has not been identified by the authors of that work [32], their experimental measurements clearly prove the ability of Physarum polycephalum to anticipate an impending environmental change. More specifically, the locomotion speed of the Physarum polycephalum was measured when favorable environmental conditions (26◦C and 90% humidity) were interrupted by three equally spaced 10 min pulses of unfavorable environmental conditions (22◦C and 60% humidity) [32]. The time separation between the pulses τ was selected between 30 and 90 minutes. It was observed that the locomotion speed at favorable conditions (approximately 0.25 mm/10min as shown in Fig. 1 of Ref. [32]) turns to close to zero each time the unfavorable conditions were presented. However, spontaneous in-phase slow downs were observed after time intervals τ, 2τ and even 3τ after the last application of unfavorable conditions. In addition, if -- after a long period of favorable conditions -- a single pulse of unfavorable condi- tions is applied again then a spontaneous slow down (called a spontaneous in-phase slow down after one disappearance [32]) after a time interval τ was observed. It clearly follows form this experiment that the Physarum polycephalum has a mecha- nism to memorize ("learn") the periodicity of environmental changes and adapts its behavior in anticipation of next changes to come [32]. We have developed a circuit model [15] of the adaptive behavior of the slime mold which has been later realized experimentally [39] using vanadium dioxide as memory element [40, 41]. The learning circuit is shown in Fig. 3(a). Here, the role of environmental conditions is played by the input voltage V (t) and the speed of loco- motion is mapped into the voltage across the capacitor C. The learning circuit design resembles a damped LC contour in which the amount of damping is controlled by the state of the memristive system M. To understand the circuit operation, we note that the memristive system employed in the circuit is of a threshold type (see Eqs. (7) and (8)), namely, its state can be significantly changed only by a voltage (across the memristive system) with a magnitude exceeding a certain threshold. Fig. 3(b) presents the switching function f (V ) used to describe a threshold-type memristive system. Our simulations of the learning circuit response to irregular and regular se- quences of pulses are shown in Fig. 4. In these simulations, the scheme described above has been used with the only restriction that the response signal cannot exceed a certain value [15] (electronically, a cut-off can be easily obtained by using an addi- tional resistor and diode). When an irregular sequence of pulses is applied to the cir- cuit, the voltage oscillations across the capacitor can not exceed the threshold volt- age of the memristive system M which continues to stay in its initial low-resistance state, thus damping the circuit. When the pulses are applied periodically with a pe- riod close to the period of the LC contour oscillations, a sufficiently strong voltage Biologically-Inspired Electronics with Memory Circuit Elements 9 Fig. 4 Modeling of the spontaneous in-phase slow-down responses [15]. This plot demonstrates that stronger and longer-lasting responses for both spontaneous in-phase slow down and sponta- neous in-phase slow down after one disappearance of the stimulus are observed only when the circuit was previously trained by a periodic sequence of three equally spaced pulses as present in V2(t). The applied voltage V1(t) is irregular and thus the three first pulses do not "train" the circuit. Reprinted with permission from Ref. [15]. c(cid:13)2009 American Physical Society. across the capacitor C is induced. This voltage switches the memristive system into the high-resistance state. Therefore, in this case, oscillations in the contour are less damped and last longer as Fig. 4 demonstrates. These oscillations exactly model the spontaneous in-phase slow down and in-phase slow down after one disappearance effects observed experimentally [32]. We note that a single learning circuit memo- rizes past events of a frequency close to the resonance frequency of LC contour. An array of learning circuits would model the learning of Physarum polycephalum in the whole frequency range [15]. Recently, an experimental implementation of this learning circuit has been re- ported [39]. In this work, a learning circuit similar to that in Fig. 3(a) has been built with the only difference that the memristive system (a vanadium dioxide memristive device [39]) has been connected in series with a capacitor (Fig. 5(a)). The memris- tive properties of vanadium dioxide are based on an insulator-to-metal phase tran- sition occurring in this material in the vicinity of 342K [40, 41]. In order to realize the memristive functionality, the vanadium dioxide device is heated to a steady-state temperature of 339.80 K (right below the transition temperature) and subjected to an externally applied voltage. The Joule heating (due to the applied voltage) incre- mentally drives the vanadium dioxide material through the phase transition, thus reducing its resistance. The operation of the learning circuit depicted in Fig. 5(a) is then clear. While off-resonance signals applied to the circuit can not excite a suffi- cient current to drive the vanadium dioxide through the phase transition, the current generated by resonance signals is sufficient for this purpose. Fig. 5(b) demonstrates modification of the transfer function of the circuit caused by off-resonance and res- 050100950100010501100-3-2-101234-3-2-101234V2(t)Voltage (arb.units) Response (arb. units)Time (s)V1(t) 10 Massimiliano Di Ventra and Yuriy V. Pershin Fig. 5 Experimental realization of the learning circuit (adaptive filter) based on vanadium dioxide memristive device. (a) Schematic of the adaptive filter in which the memristive device (with a small memcapacitive component) is connected in series with a capacitor C and inductor L. We note that such realization of the learning circuit operates similarly to the learning circuit shown in Fig. 3(a). (b) Small-signal (10 mV) transfer function (Vout /Vin) for the adaptive filter plotted before and after off-resonance "A" and on-resonance "B" pulses. Solid lines are RLC bandpass-filter fit to data, which generates the ω0 and Q values in the legend. Pulse sequence B has a significant training effect on the circuit, while A has little or no effect. (c) Time series of the off-resonance "A" sequence of pulses and on-resonance "B" sequence of pulses. Reprinted with permission from Ref. [39]. c(cid:13)2010 American Institute of Physics. onance pulse sequences (Fig. 5(c)) applied to the circuit. Fig. 5(b) clearly indicates a change in the transfer function caused by resonance signals (learning). Finally, we mention that the formalism of memory circuit elements [4] has also been useful in modeling biophysical systems whose electric response depends on the history of applied voltages or currents. An example of such situation is the electro- osmosis in skin which has been recently described by a memristive model [42]. Physically, the voltage applied to the skin induces a water flow in sweat-duct cap- illaries changing the skin conductance. The position of the water table (the level separating dry and wet zones) in capillaries plays the role of the internal state vari- able whose dynamics is determined by the applied voltage [42]. The memristive model of electro-osmosis [42] is in a good agreement with experimental data and further demonstrates the potential of the formalism of memory circuit elements for modeling biophysical processes. 4.2 Neuromorphic circuits Our second example of biologically-inspired circuits with memory circuit elements is from the area of neural networks. Neural networks form a class of circuits whose Biologically-Inspired Electronics with Memory Circuit Elements 11 Fig. 6 Memristive neural network with an associative memory ability. Here, two input neurons (N1 and N2) are connected through mem- ristive synapses (S1 and S2) to the output neuron N3. The details of circuit operation are given in the text. Reprinted from Ref. [44]. c(cid:13)2010 with permission from Elsevier. operation mimics the operation of the human (and animal) brain. Below, we con- sider electronic implementations of two important processes occurring in biolog- ical neural networks: associative memory and spike timing-dependent plasticity. Both features can be implemented in artificial neural networks based on memris- tive synapses. 4.2.1 Associative memory The associative memory is one of the most fundamental functionalities of the hu- man (and animal) brain. By making associations we learn, adapt to a changing en- vironment and better retain and recall events. One of the most famous experiments related to associative memory is Pavlov's experiment [43] whereby salivation of a dog's mouth is first set by the sight of food. Then, if the sight of food is accompanied by a sound (e.g., the tone of a bell) over a certain period of time, the dog learns to associate the sound to the food, and salivation can be triggered by the sound alone, without the intervention of vision. Recently, we have reproduced [44] the Pavlov's experiment utilizing a neural network with memristive synapses. As a first example, we have implemented the well known Hebbian rule introduced by Hebb in 1949: "when an axon of cell A is near enough to excite a cell B and repeatedly or persistently takes part in firing it, some growth process or metabolic change takes place in one or both cells such that A's efficiency, as one of the cells firing B, is increased" [45]. To put it differently, the neurons that fire together, wire together. In order to show associative memory, let us consider a simple neural network consisting of three electronic neurons and two memristive synapses as shown in Fig. 6. We assume that the first input neuron activates under a specific ("visual") event, such as "sight of food", and the second input neuron activates under another ("auditory") event, such as a particular "sound". On the electronic level, an electronic neuron sends forward (to its output) and backward (to its input) voltage spikes of opposite polarity when the amplitude of the input signal exceeds a threshold value. Regarding the dynamics of memris- tive synapses, they have been selected of a threshold-type (Eqs. (7) and (8)) with a threshold voltage exceeding the output voltage of electronic neurons. In this case, N1S1Input 1"ihffd"1NS1Output"sight of food"N3SInput 2"salivation"N2S2p"sound" 12 Massimiliano Di Ventra and Yuriy V. Pershin voltage spikes applied to a single terminal of a memristive synapse is not enough to induce its change. The latter is possible only if forward and backward propagat- ing spikes overlap in time across a synapse. We have employed memristor emula- tors [46, 44] as memristive synapses2. The main components of a memristor emula- tor are a digital potentiometer, a microcontroller and an analog-to-digital converter. Using the converter, the microcontroller continuously reads the voltage applied to the digital potentiometer and updates the potentiometer resistance according to a pre-programmed model of a voltage- or current-controlled memristive system. The operation of electronic neurons is realized along similar lines [44]. Operation of the associative memory is presented in Fig. 7 where a detail of this process is given. Our work as described above [44] demonstrates the potential of memristive de- vices for neuromorphic circuits applications. Importantly, it has been recently shown in numerous experiments that memristive devices can be built at the nanoscale [52, 53, 54, 55, 56, 57, 14, 58, 59, 18, 60, 6, 61, 62]. This opens up the possibility to Fig. 7 Experimental demonstration of the associative memory with memristive neural networks. In this experiment, a simple neural network shown in Fig. 6 was realized. The first "probing" phase demonstrates that, initially, only a signal from N1 neuron activates the output neuron. The association of the Input 2 signal with the Output develops in the "learning phase" when N1 and N2 neurons are simultaneously activated. In this case, a signal at the Input 1 excites the third neuron that sends back-propagating pulses of a negative amplitude. These pulses, when applied simultaneously with forward propagating pulses from the Input 2 to the second memristive synapse S2 cause it to learn. The final "probing" phase demonstrates that signals from both N1 and N2 activate the output neuron. A detalied description of the experiment is given in Ref. [44]. Reprinted from Ref. [44]. c(cid:13)2010 with permission from Elsevier. 2 Several designs of memristor [46, 47, 48, 7] as well as memcapacitor and meminductor [47, 49, 50, 51] emulators are known in the literature. These emulators serve as an important practical tool to build small-scale circuits with memory circuit elements. 03691215-202468101214ProbingOutput"salivation"Input 2"sound"LearningProbingInput 1"sight of food"Voltage (V)Time (s) Biologically-Inspired Electronics with Memory Circuit Elements 13 fabricate neuromorphic circuits with the same amount of memristive synapses as the number of biological synapses in the human brain (∼ 1014). In fact, one of the main challenges for practical realizations of an artificial brain on a chip is related to the high connectivity of biological neurons. It has been estimated that, on average, the number of connections per neuron is of the order of 103. Therefore, neural net- works of memelements built at the nanoscale offer advantages - in terms of density - unavailable with current active elements (such as transistors). 4.2.2 Spike timing-dependent plasticity However, the above mentioned simple Hebbian rule does not describe the much more complicated time-dependent plasticity of biological synapses [63, 64, 65, 66]. The latter has come to be known as spike timing-dependent plasticity (STDP). When a post-synaptic signal reaches the synapse before the action potential of the pre- synaptic neuron, the synapse shows long-term depression (LTD), namely its strength decreases (smaller connection between the neurons) depending on the time differ- ence between the post-synaptic and the pre-synaptic signals. Conversely, when the post-synaptic action potential reaches the synapse after the pre-synaptic action po- tential, the synapse undergoes a long-time potentiation (LTP), namely the signal transmission between the two neurons increases in proportion to the time difference between the pre-synaptic and the post-synaptic signals. The learning process and the storing of information in the brain thus follow non-trivial time-dependent features which have not been fully understood yet. Implementation of STDP in artificial net- works can thus help unraveling these mechanisms. The spike timing-dependent plasticity can be implemented using different types of memristive systems. Following our previous work [67], neuromorphic circuits can be based on memristive systems with or without an internal spike-timing track- ing capability. In the most simple case, memristive systems without spike-timing tracking capability are of the first order, while those supporting such capability are of the second order as an additional internal state variable is needed to track timing of pre-synaptic and post-synaptic pulses [67]. In the first case, an additional exter- nal hardware is required to implement the spike timing-dependent plasticity. For example, STDP was recently realized using a combination of memristive systems with CMOS (complementary metal-oxide-semiconductor) elements [60] (see Fig. 8). Another approach involves utilization of overlapping pulses of opposite polari- ties [67, 68, 69]. A simple second-order memristive system with time-tracking capability is de- scribed by the following equations [67] R = x, x = γ [θ (VM −Vt )θ (y− yt ) + θ (−VM −Vt )θ (−y− yt )]y, y = τ [−VMθ (VM −Vt )θ (yt − y)−VMθ (−VM −Vt )θ (y + yt )− y] , 1 (9) (10) (11) 14 Massimiliano Di Ventra and Yuriy V. Pershin Fig. 8 (a) Measured change in the synaptic weight versus spike separation. Inset: SEM image of the memristive crossbar array, scale bar is 300 nm. (b) Measured change in excitatory postsynaptic current of rat neurons after repetitive correlated spiking versus relative spiking timing (the plot was reconstructed from Ref. [65]). Inset: image of a hippocampal neuron (the image was adapted with permission from reference [70]). Scale bar is 50 µm. Reprinted with permission from [60]. Copyright 2010 American Chemical Society. where x and y are internal state variables, γ is a constant, Vt is a threshold voltage, yt is the threshold value of y, and τ is a constant defining the time window of STDP. The second-order memristive system with timing tracking capability defined by the above equations is very promising for neuromorphic circuits application since neu- ron's firing can be implemented simply by short single rectangular pulses and no ad- ditional hardware as in the case of first-order memristors (see, e.g., Ref. [60]). How- ever, such solid-state second-order memristors need to be developed, even though their implementation in memristor emulators [46, 47] can be easily realized. Moreover, several authors have discussed applications of three-terminal transistor- like electronic devices with memory [25, 26, 71] in the area of neuromorphic com- puting. Although, formally such devices can not be categorized as memristive sys- tems, their operation is clearly based on memristive effects. In particular, Lai et al. [26] have experimentally fabricated a synaptic transistor. For instance, Fig. 9 depicts their experimental scheme and selected measurement results that confirm realization of spike timing-dependent plasticity in this device. Biologically-Inspired Electronics with Memory Circuit Elements 15 (a) (b) Fig. 9 (a) Structure of synaptic transistor . (b) The relative changes of the postsynaptic currents measured after application of 120 pairs of temporally correlated pre- and post-synaptic spikes. From [26], copyright Wiley-VCH Verlag GmbH & Co. KGaA. Reproduced with permission. 4.3 Networks of memory circuit elements A human brain - and also the brain of many other living organisms - solves many problems much better than traditional computers. The reason for this is a type of massive parallelism in which a large number of neurons and synapses participate simultaneously in the calculation. Here, we consider networks of memory circuit elements and their ability to i) solve efficiently certain graph theory optimization problems, and ii) retain such information for later use. In particular, we demonstrate that a network of memristive devices solves the maze problem much faster than any existing algorithm [72]. Similar to the brain's operation, such an extraordinary ad- vance in computational power is due to the massively-parallel network dynamics in which all network components are simultaneously involved in the calculation. This type of parallelism could be dubbed as an analog parallelism which is very differ- ent from that used in conventional supercomputers. In the latter systems, each core typically runs a separate process that, relatively rarely, exchanges information with other cores. In calculations done by networks, the information exchange is contin- uous resulting in a tremendous increase of computational power as we demonstrate below. Left panel of Fig. 10 depicts a memristive network (memristive processor) in which points of a square grid are connected by basic units (memristive system plus switch (FET)) [72]. Each switch in the network can be in the "connected" or "not- connected" state. Since the direction of current flow in the network is not known a priori, the polarity of adjacent memristive devices (indicated by the black thick line in the memristor symbol in Figure 1) is chosen to be alternating. Experimentally, the suggested network could be fabricated using, e.g., CMOL (Cmos+MOLecular-scale devices) architecture [73] combining a single memristor layer with a conventional CMOS layer. The operation of the massively-parallel processor consists of three 16 Massimiliano Di Ventra and Yuriy V. Pershin Fig. 10 Maze mapping into a network of memristors. Right panel. The maze is covered by an array of vertical and horizontal lines having the periodicity of the maze. Left panel. Architecture of the network of memristors in which each crossing between vertical and horizontal lines in the array (in the right panel) is represented by a grid point to which several basic units consisting of memristors and switches (field-effect transistors) are linked in series. The maze topology is encoded into the state of the switches such that if the short line segment connecting neighboring crossing points in the array crosses the maze wall then the state of the corresponding switch is "not connected" (shown with red symbols). All other switches are in the "connected" state. The external voltage (V ) is applied across the connection points corresponding to the entrance (V ) and exit (ground, GND) points of the maze. Reprinted with permission from Ref. [72]. c(cid:13)2011 American Physical Society. main stages: initialization, computation and reading out the computation result. All these stages are realized by externally applied signals (originating, e.g., from the CMOS layer). During the first initialization stage, all memristive elements in the network are switched into the "OFF" state. This can be done, for example, by applying GND and appropriately selected V1 voltages in a chessboard-like pattern to all grid points of the memristive network for a sufficiently long period of time [72]. After that, the maze topology is mapped onto the memristive network by setting appropriate switches into the "not connected" state. We describe this process in the caption of Fig. 10. The computation stage consists in the application of a single voltage pulse of appropriate amplitude and duration across grid points corresponding to the entrance and exit points of the maze. The solution can be later read or used in further calculations. We have modeled the memristive processor operation by numerically solving Kirchhoff's current law equations complemented by Eqs. (5), (6) which in the present network case are modified as Biologically-Inspired Electronics with Memory Circuit Elements 17 Fig. 11 Solution of a multiple-path maze [72]. The maze solution contains two common segments (red dots connected by a red line), and two alternative segments of different lengths close to the left bottom corner. The memristance in the shorter segment (blue dots connected by a red line) is smaller than that in the longer segment (green dots connected by a red line) since the current through the shorter segment is larger and, consequently, the change of the memristors' state along this segment is larger. The arrow at the bottom indicates a splitting point of the solution path. The resistance is in Ohms, the voltage is in Volts and the current is in Amperes. Reprinted with permission from Ref. [72]. c(cid:13)2011 American Physical Society. i j = RONxi j + ROFF (1− xi j) , RM (12) where RON and ROFF are again the minimal and maximal values of memristance, xi j is the dimensionless internal state variable for each memristor bound to the region 0 ≤ xi j ≤ 1, and (i, j) are grid indexes of a memristor to identify its location in the network. The dynamics of xi j is then given by dxi j dt = αIi j(t), (13) with α a constant and Ii j(t) the current flowing through the memristor (i j). Fig. 11(a) shows a solution of a multiple-path maze. The maze solution is clearly seen in Fig. 11 as chains of red, blue and green boxes (representing memristive devices with lower memristance) connected by a red line. Importantly, the memris- tive processor not only determines all possible solutions of the maze but also stores them and sorts them out according to their length. This feature is described in more details in the caption of Fig. 11. Also, the memristive processor requires only one single step to find the maze solution thus outperforming all known maze solving approaches and algorithms. 18 Massimiliano Di Ventra and Yuriy V. Pershin We also note that the wide selection of physical mechanisms of memory we can "shop" from, offers many opportunities to design novel efficient electronic de- vices [6]. For example, a memristive processor based on fast switching nanoionic metal/insulator/metal cells [6] would require just few nanoseconds or even less 3 to solve the maze. More generally, a network of memristors - or other memory cir- cuit elements - can be considered as an adaptable medium whose state dynamically changes in response to time-dependent signals or changes in the network config- uration. Therefore, the use of these processors is not limited to maze solving: We expect they could help find the solution of many graph theory optimization prob- lems including, for instance, the traveling salesman problem, shortest path problem, etc. 5 Conclusions and Outlook In conclusion, we have shown that the two-terminal electronic devices with memory -- memristive, memcapacitive and meminductive systems -- are very useful to model a variety of biological processes and systems. The electronic implementation of all these mechanisms can clearly lead to a novel generation of "smart" electronic circuits that can find useful applications in diverse areas of science and technology. In addition, these memelements and their networks, provide solid ground to test various hypothesis and ideas regarding the functioning of the human (and animal) brain both theoretically and experimentally. Theoretically because their flexibility in terms of what type and how many internal state variables responsible for memory, or what network topology are required to reproduce certain biological functions can lead to a better understanding of the microscopic mechanisms that are responsible for such features in living organisms. Experimentally because with the continuing miniaturization of electronic devices, memelements can be assembled into networks with similar densities as the biological systems (e.g., the brain) they are designed to emulate. In particular, we anticipate potential applications for memcapacitive and meminductive systems [4] which offer such an important property as low energy dissipation combined with information storage capabilities. We are thus confident that the area of biologically-inspired electronics with memory circuit elements will offer many research opportunities in several fields of science and technology. Acknowledgements M.D. acknowledges partial support from the NSF Grant No. DMR-0802830. 3 Fast sub-nanosecond switching has been recently reported in tantalum oxide memristive sys- tems [62]. Biologically-Inspired Electronics with Memory Circuit Elements References 19 1. S.J. Russell, P. Norvig, Artificial Intelligence: A Modern Approach, 3rd edn. (Prentice Hall, 2009) 2. A. Colorni, M. M. Dorigo, V. Maniezzo, in Actes de la premire confrence europenne sur la vie artificielle, Paris, France, Elsevier Publishing (1991), pp. 134 -- 142 3. N. Monmarche, F. Guinand, P. Siarry, Artificial Ants (Wiley-ISTE, 2010) 4. M. Di Ventra, Y.V. Pershin, L.O. Chua, Proc. IEEE 97(10), 1717 (2009) 5. M. Di Ventra, Y.V. Pershin, Mat. Today 14, 584 (2011) 6. Y.V. Pershin, M. Di Ventra, Advances in Physics 60, 145 (2011) 7. L.O. Chua, IEEE Trans. Circuit Theory 18, 507 (1971) 8. L.O. Chua, S.M. Kang, Proc. IEEE 64, 209 (1976) 9. Y.V. Pershin, M. Di Ventra, Phys. Rev. B 78, 113309 (2008) 10. X. Wang, Y. Chen, H. Xi, H. Li, D. Dimitrov, El. Dev. Lett. 30, 294 (2009) 11. J. Martinez-Rincon, Y.V. Pershin, IEEE Trans. El. Dev. 58, 1809 (2011) 12. J. Martinez-Rincon, M. Di Ventra, Y.V. Pershin, Phys. Rev. B 81, 195430 (2010) 13. F. Corinto, A. Ascoli, M. Gilli, submitted for publication (2011) 14. D.B. Strukov, G.S. Snider, D.R. Stewart, R.S. Williams, Nature 453, 80 (2008) 15. Y.V. Pershin, S. La Fontaine, M. Di Ventra, Phys. Rev. E 80, 021926 (2009) 16. M. Krems, Y.V. Pershin, M. Di Ventra, Nano Lett. 10, 2674 (2010) 17. A. Stotland, M. Di Ventra, arXiv:1104.4485v2 (2011) 18. C. Schindler, G. Staikov, R. Waser, Appl. Phys. Lett. 94, 072109 (2009) 19. Q. Lai, L. Zhang, Z. Li, W.F. Stickle, R.S. Williams, Y. Chen, Appl. Phys. Lett. 95, 213503 (2009) Tech. 49(11), 2093 (2001) 20. V. Lubecke, B. Barber, E. Chan, D. Lopez, M. Gross, P. Gammel, IEEE Trans. Microw. Theory 21. I. Zine-El-Abidine, M. Okoniewski, J.G. McRory, in Proceedings of the 2004 International Conference on MEMS, NANO and Smart Systems (ICMENS'04) (2004), pp. 636 -- 638 22. S. Chang, S. Sivoththaman, IEEE El. Dev. Lett. 27(11), 905 (2006) 23. S. Liu, N. Wu, A. Ignatiev, J. Li, J. Appl. Phys. 100, 056101 (2006) 24. V.V. Erokhin, T.S. Berzina, M.P. Fontana, Crystallogr. Rep. 52, 159 (2007) 25. F. Alibart, S. Pleutin, D. Guerin, C. Novembre, S. Lenfant, K. Lmimouni, C. Gamrat, D. Vuil- laume, Adv. Funct. Mat. 20, 330 (2010) 26. Q. Lai, L. Zhang, Z. Li, W.F. Stickle, R.S. Williams, Y. Chen, Adv. Mat. 22, 2448 (2010) 27. T. Kozlowski, S. Pallardy, Botanical Review 68, 270 (2002) 28. A. Trewavas, Annals of Botany 92, 1 (1998) 29. P. Calvo Garzon, F. Keijzer, Adaptive Behavior 19, 155 (2011) 30. J.E.R. Staddon, Adaptive behavior and learning (Cambridge University Press, 1983) 31. J. Ojal, Biol. Cybern. 79, 403 (1998) 32. T. Saigusa, A. Tero, T. Nakagaki, Y. Kuramoto, Phys. Rev. Lett. 100, 018101 (2008) 33. Q. Li, B. McNeil, L.M. Harvey, Free Radical Biology and Medicine 44, 394 (2008) 34. J. Hartley, J.W.G. Cairney, A.A. Meharg, Plant and Soil 189, 303 (1997) 35. H. Thieringer, P. Jones, M. Inouye, BioEssays 20, 49 (1998) 36. J. van der Oost, M.M. Jore, E.R. Westra, M. Lundgren, S.J.J. Brouns, Trends in Biochemical Sciences 34, 401 (2009) 37. J. Larsson, J.A.A. Nylander, B. Bergman, BMC evolutionary biology 11, 187 (2011) 38. J.H. Holland, Adaptation in Natural and Artificial Systems: An Introductory Analysis with Applications to Biology, Control, and Artificial Intelligence (Bradford Book, 1992) 39. T. Driscoll, J. Quinn, S. Klein, H.T. Kim, B.J. Kim, Y.V. Pershin, M. Di Ventra, D.N. Basov, Appl. Phys. Lett. 97, 093502 (2010) 40. T. Driscoll, H.T. Kim, B.G. Chae, M. Di Ventra, D.N. Basov, Appl. Phys. Lett. 95, 043503 (2009) 41. T. Driscoll, H.T. Kim, B.G. Chae, B.J. Kim, Y.W. Lee, N.M. Jokerst, S. Palit, D.R. Smith, M. Di Ventra, D.N. Basov, Science 325, 1518 (2009) 20 Massimiliano Di Ventra and Yuriy V. Pershin 42. G.K. Johnsen, C.A. Lutken, O.G. Martinsen, S. Grimnes, Phys. Rev. E 83, 031916 (2011) 43. I. Pavlov, Conditioned Reflexes: An Investigation of the Physiological Activity of the Cerebral Cortex (London: Oxford University Press, 1927). (translated by G. V. Anrep) 44. Y.V. Pershin, M. Di Ventra, Neural Networks 23, 881 (2010) 45. D.O. Hebb, The Organization of Behavior; A Neuropsychological Theory (New York: Wiley, 1949) 46. Y.V. Pershin, M. Di Ventra, IEEE Trans. Circ. Syst. I 57, 1857 (2010) 47. D. Biolek, V. Biolkova, Int. J. Numer. Mod. (in press) (2011) 48. D. Biolek, V. Biolkova, Z. Kolka, in 5th International Conference on Circuits, Systems and Signals (CSS'11). (2011), p. 171 49. D. Biolek, V. Biolkova, El. Lett. 46, 1428 (2010) 50. Y.V. Pershin, M. Di Ventra, El. Lett. 47, 243 (2011) 51. Y.V. Pershin, M. Di Ventra, El. Lett. 46, 517 (2010) 52. T. Tamura, T. Hasegawa, K. Terabe, T. Nakayama, T. Sakamoto, H. Sunamura, H. Kawaura, S. Hosaka, M. Aono, Jpn. J. Appl. Phys. 45, L364 (2006) 53. R.R. Waser, M. Aono, Nat. Mat. 6, 833 (2007) 54. D. Lee, D. jun Seong, I. Jo, F. Xiang, R. Dong, S. Oh, , H. Hwang, Appl. Phys. Lett. 90, 122104 (2007) 55. S. Dietrich, M. Angerbauer, M. Ivanov, D. Gogl, H. Hoenigschmid, M. Kund, C. Liaw, M. Markert, IEEE J. Sol.-State Circ. 42, 839 (2007) 56. S. Raoux, G.W. Burr, M.J. Breitwisch, C.T. Rettner, Y.C. Chen, R.M. Shelby, M. Salinga, D. Krebs, S.H. Chen, H.L. Lung, C.H. Lam, IBM J. Res. Dev. 52, 465 (2008) 57. I.H. Inoue, S. Yasuda, H. Akinaga, H. Takagi, Phys. Rev. B 77, 035105 (2008) 58. A. Sawa, Mat. Today 11, 28 (2008) 59. S.H. Jo, K.H. Kim, W. Lu, Nano Lett. 9, 870 (2009) 60. S.H. Jo, T. Chang, I. Ebong, B.B. Bhadviya, P. Mazumder, W. Lu, Nano Lett. 10, 1297 (2010) 61. M.J. Lee, C.B. Lee, D. Lee, S.R. Lee, M. Chang, J.H. Hur, Y.B. Kim, C.J. Kim, D.H. Seo, S. Seo, U.I. Chung, I.K. Yoo, K. Kim, Nature Materials 10(8), 625 (2011) 62. A.C. Torrezan, J.P. Strachan, G. Medeiros-Ribeiro, R.S. Williams, Nanotechnology 22(48), 485203 (2011) 63. W.B. Levy, O. Steward, Neuroscience 8, 791 (1983) 64. H. Markram, J. Lubke, M. Frotscher, B. Sakmann, Science 275, 213 (1997) 65. G.Q. Bi, M.M. Poo, J. Neurosci. 18, 10464 (1998) 66. R.C. Froemke, Y. Dan, Nature 416, 433 (2002) 67. Y.V. Pershin, M. Di Ventra, Proc. IEEE (in press); arXive:1009.6025 (2010) 68. G.S. Snider, SciDAC Review 10, 58 (2008) 69. S. Parkin, Innately three dimensional spintronic memory and logic devices: Racetrack memory and spin synapses (2010). Talk at 2010 MRS SprinG Meeting 70. S. Kaech, G. Banker, Nat Protoc. 1, 2406 (2006) 71. W.S. Zhao, G. Agnus, V. Derycke, A. Filoramo, J.P. Bourgoin, C. Gamrat, Nanotechnology 21(17), 175202 (2010) 72. Y.V. Pershin, M. Di Ventra, Phys. Rev. E 84, 046703 (2011) 73. K.K. Likharev, D.B. Strukov, in Introducing Molecular Electronics, vol. 657, ed. by G.F. G. Cuniberti, K. Richter (Springer, 2005), pp. 447 -- 477
1305.4160
1
1305
2013-05-17T19:17:38
A generative spike train model with time-structured higher order correlations
[ "q-bio.NC", "math.DS" ]
Emerging technologies are revealing the spiking activity in ever larger neural ensembles. Frequently, this spiking is far from independent, with correlations in the spike times of different cells. Understanding how such correlations impact the dynamics and function of neural ensembles remains an important open problem. Here we describe a new, generative model for correlated spike trains that can exhibit many of the features observed in data. Extending prior work in mathematical finance, this generalized thinning and shift (GTaS) model creates marginally Poisson spike trains with diverse temporal correlation structures. We give several examples which highlight the model's flexibility and utility. For instance, we use it to examine how a neural network responds to highly structured patterns of inputs. We then show that the GTaS model is analytically tractable, and derive cumulant densities of all orders in terms of model parameters. The GTaS framework can therefore be an important tool in the experimental and theoretical exploration of neural dynamics.
q-bio.NC
q-bio
A generative spike train model with time-structured higher order correlations James Trousdale, Yu Hu, Eric Shea-Brown and Kresimir Josi´c Abstract Emerging technologies are revealing the spiking activity in ever larger neural ensembles. Frequently, this spiking is far from independent, with correlations in the spike times of different cells. Under- standing how such correlations impact the dynamics and function of neural ensembles remains an important open problem. Here we describe a new, generative model for correlated spike trains that can exhibit many of the features observed in data. Extending prior work in mathematical finance, this generalized thinning and shift (GTaS) model creates marginally Poisson spike trains with diverse temporal correlation structures. We give several examples which highlight the model's flexibility and utility. For instance, we use it to examine how a neural network responds to highly structured patterns of inputs. We then show that the GTaS model is analytically tractable, and derive cumulant densities of all orders in terms of model parameters. The GTaS framework can therefore be an important tool in the experimental and theoretical exploration of neural dynamics. 1 Introduction Recordings across the brain suggest that neural populations spike collectively -- the statistics of their activity as a group are distinct from that expected in assembling the spikes from one cell at a time [6, 48, 64, 62, 67, 8, 31, 32, 10, 24, 57]. Advances in electrode and imaging technology allow us to explore the dynamics of neural populations by simultaneously recording the activity of hundreds of cells. This is revealing patterns of collective spiking that extend across multiple cells. The underlying structure is intriguing: For example, higher-order interactions among cell groups have been observed widely [67, 68, 4, 56, 48, 64, 76, 24]. A number of recent studies point to mechanisms that generate such higher-order correlations from common input processes, including unobserved neurons. This suggests that, in a given recording or given set of neurons projecting downstream, higher-order correlations may be quite ubiquitous [51, 79, 44, 9]. Moreover, these higher-order correlations may impact encoded information [55, 17, 24] as well as the firing rate of downstream neurons [46]. What exactly is the impact of such collective spiking on the encoding and transmission of information in the brain? This question has been studied extensively, but much remains unknown. Results to date show that the answers will be varied and rich. Patterned spiking can impact responses at the level of single cells [78, 62, 46] and neural populations [59, 58, 74, 5]. Neurons with even the simplest of nonlinearities can be highly sensitive to correlations in their inputs. Moreover, such nonlinearities are sufficient to accurately decode signals from the input to correlated neural populations [66]. 1 An essential tool in understanding the impact of collective spiking is the ability to generate artificial spike trains with a predetermined structure across cells and across time [29, 45, 50, 14]. Such synthetic spike trains are the grist for testing hypotheses about spatiotemporal patterns in coding and dynamics. In experimental studies, such spike trains can be used to provide structured stimulation of single cells across their dendritic trees via glutamate uncaging [23, 26, 12, 13], or entire populations of neurons via optical stimulation of microbial opsins [19, 30]. Computationally, they are used to examine the response of nonlinear models of downstream cells [62, 46, 18]. Therefore, much effort has been devoted to developing statistical models of population activity. A number of flexible, yet tractable probabilistic models of joint neuronal activity have been pro- posed. Pairwise correlations are the most common type of interactions obtained from multi-unit recordings. Therefore many earlier models were designed to generate samples of neural activity patterns with predetermined first and second order statistics [29, 45, 50, 14]. In these models, higher-order correlations are not explicitly and separately controlled. A number of different models have been used to analyze higher-order interactions. However, most of these models assume that interactions between different cells are instantaneous (or near- instantaneous) [70, 46, 40]. A notable exception is the work of [11], which developed such methods for use in financial applications. In these previous efforts, correlations at all orders were character- ized by the increase, or decrease, in the probability that groups of cells spike together at the same time, or have a common temporal correlation structure regardless of the group. The aim of the present work is to provide a statistical method for generating spike trains with more general correlation structures across cells and time. Specifically, we allow distinct temporal structure for correlations at pairwise, triplet, and all higher orders, and do so separately for different groups of cells in the neural population. Our aim to describe a model that can be applied in neuroscience, and can potentially be fit to emerging datasets. A sample processes from our model is shown in Fig. 1. The multivariate spike train consists of six marginally Poisson processes. Each event was either uncorrelated with all other events across the population, or correlated in time with an event in all other spike trains. This model was configured to exhibit activity that cascades through a sequence of neurons. Specifically, neurons with larger index tend to fire later in a population wide event (this is similar to a synfire chain [2], but with variable timing of spikes within the cascade). In Fig. 1B, we plot the "population cross-cumulant density" for three chosen neurons -- the summed activity of the population triggered by a spike in a chosen cell. The center of mass of this function measures the average latency by which spikes of the neuron in question precede those of the rest of the population [48]. Finally, Fig. 1C shows the third-order cross-cumulant density for the three neurons. The triangular support of this function is a reflection of a synfire-like cascade structure of the spiking shown in the raster plot of panel A: when firing events are correlated between trains, they tend to proceed in order of increasing index. We demonstrate the impact of such structured activity on a downstream network in Section 2.2. 2 Results Our aim is to describe a flexible multivariate point process capable of generating a range of high order correlation structures. To do so we extend the TaS (thinning and shift) model of temporally- and spatially-correlated, marginally Poisson counting processes [11]. The TaS model itself gener- alizes the SIP and MIP models [46] which have been used in theoretical neuroscience [17, 59, 73]. However the TaS model has not been used as widely. The original TaS model is too rigid to generate a number of interesting activity patterns observed in multi-unit recordings [37, 49, 48]. 2 Figure 1: (A) Raster plot of event times for an example multivariate Poisson process X = (X1, . . . , X6) generated using the methods presented below. This model exhibits independent marginal events (blue) and population-level, chain-like events (red). (B) Some second order pop- ulation cumulant densities (i.e., second order correlation between individual unit activities and population activity) for this model [48]. Greater mass to the right (resp. left) of τ = 0 indi- follow) in pairwise-correlated events. (C) Third-order cates that the cell tends to lead (resp. cross-cumulant density for processes X1, X2, X3. The quantity κX 123(τ1, τ2) yields the probability of observing spikes in cells 2 and 3 at an offset τ1, τ2 from a spike in cell 1, respectively, in excess of what would be predicted from the first and second order cumulant structure. All quantities are precisely defined in the Methods. Note: system parameters necessary to reproduce results are given in the Appendix for all figures. We therefore developed the generalized thinning and shift model (GTaS) which allows for a more diverse temporal correlation structure. We begin by describing the algorithm for sampling from the GTaS model. This constructive approach provides an intuitive understanding of the model's properties. We then present a pair of examples, the first of which highlights the utility of the GTaS framework. The second example demonstrates how sample point processes from the TaS models can be used to study population dynamics. Next, we present the analysis which yields the explicit forms for the cross-cumulant densities derived in the context of the examples. We do so by first establishing a useful distribu- tional representation for the GTaS process, paralleling [11]. Using this representation, we derive cross-cumulants of a GTaS counting process, as well as explicit expressions for the cross-cumulant 3 densities. After explaining the derivation at lower orders, we present a theorem which describes cross-cumulant densities at all orders. 2.1 GTaS model simulation D The GTaS model is parameterized first by a rate λ which determines the intensity of a "mother process" - a Poisson process on R. The events of the mother process are marked, and the markings determine how each event is distributed among a collection of N daughter processes. The daughter processes are indexed by the set D = {1, . . . , N}, and the set of possible markings is the power , the set of all subsets D. We define a probability distribution p = (pD)D⊂D, assigning a set 2 probability to each possible marking, D. As we will see, pD determines the probability of a joint event in all daughter processes with indices in the set D. Finally, to each marking, D, we assign a probability distribution QD, giving a family of shift (jitter) distributions (QD)D⊂D. Each (QD) is a distribution over RN . The rate λ, the distribution p over the markings, and the family of jitter distributions (QD)D⊂D, define a vector X = (X1, . . . , XN ) of dependent daughter Poisson processes described by the fol- lowing algorithm, which yields a single realization (see Fig. 2): 1. Simulate the mother Poisson process of rate λ on R, generating a sequence of event times {tj}. (Fig. 2A) 2. With probability pDj assign the subset Dj ⊂ D to the event of the mother process at time tj. This event will be assigned only to processes with indices in Dj. (Fig. 2B) 3. Generate a vector (Y j 1 , . . . , Y j N ) = Yj from the distribution QDj . For each i ∈ D, the time tj + Y j i is set as an event time for the marginal counting process Xi. (Fig. 2C) Hence copies of each point of the mother process are placed into daughter processes after a shift in time. A primary difference between the GTaS model and the TaS model presented in [11] is the dependence of the shift distributions QD on the chosen marking. This allows for greater flexibility in setting the temporal cumulant structure. 2.2 Examples Relation to SIP/MIP processes Two simple models of correlated, jointly Poisson processes were defined in [46]. The resulting spike trains exhibit spatial correlations, but only instanta- neous temporal dependencies. Each model was constructed by starting with independent Poisson processes, and applying one of two elementary point process operations: superposition and thin- ning [21]. We show that both models are special cases of the GTaS model. In the single interaction process (SIP), each marginal process Xi is obtained by merging an independent Poisson process with a common, global Poisson process. That is, tively. An SIP model is equivalent to a GTaS model with mother process rate λ = λc +(cid:80)N where Zc and each Zi are independent Poisson counting processes on R with rates λc, λi, respec- i=1 λi, i = 1, . . . , N, Xi(·) = Zi(·) + Zc(·), and marking probabilities  λi λ D = {i} λ D = D 0 otherwise λc . pD = 4 Figure 2: An illustration of a GTaS simulation. (A) Step 1: Simulate the mother process - a time-homogeneous Poisson process with event times {tj}. (B) Step 2: For each tj in step 1, select a set Dj ⊂ D according to the distribution pD, and project the event at time tj to the subsets with indices in Dj. The legend indicates the colors assigned to three possible markings in this example. (C) Step 3: For each pair (tj, Dj) generated in the previous two steps, draw Yj from QDj , and shift the event times in the daughter processes by the corresponding values Y j i . coordinate (i.e., qD(y1, . . . , yN ) ≡(cid:81)N Note that if λc = 0, each spike will be assigned to a different process Xi, resulting in N independent Poisson processes. Lastly, each shift distribution is equal to a delta distribution at zero in every i=1 δ(yi) for every D ⊂ D). Thus, all joint cumulants (among distinct marginal processes) of orders 2 through d are delta functions of equal magnitude, λpD. The multiple interaction process (MIP) consists of N Poisson processes obtained from a common mother process with rate λm by thinning [21]. The ith daughter process is formed by independent (across coordinates and events) deletion of events from the mother process with probability p = (1 − ). Hence, an event is common to k daughter processes with probability k. Therefore, if we take the perspective of retaining, rather than deleting events, the MIP model is equivalent to a GTaS process with λ = λm, and pD = D(1− )d−D. As in the SIP case, the shift distributions are singular in every coordinate. Below, we present a general result (Theorem 1) which immediately yields as a corollary that the MIP model has cross-cumulant functions which are δ functions in all dimensions, scaled by k, where k is the order of the cross-cumulant. Generation of synfire-like cascade activity The GTaS framework provides a simple, tractable way of generating cascading activity where cells fire in a preferred order across the population -- as in a synfire chain, but (in general) with variable timing of spikes [2, 37, 3, 1, 7]. More generally, it can be used to simulate the activity of cell assemblies [32, 34, 16, 10], in which the firing of groups of neurons is likely to follow a particular order. In the Introduction, we briefly presented one example in which the GTaS framework was used to generate synfire-like cascade activity (see Fig. 1), and we present another in Fig. 3. In what follows, we will present the explicit definition of this second model, and then derive explicit expressions for its cumulant structure. Our aim is to illustrate the diverse range of possible correlation structures that can be generated using the GTaS model. Consider an N -dimensional counting process X = (X1, . . . , XN ) of GTaS type, where N ≥ 4. We restrict the marking distribution so that pD ≡ 0 unless D ≤ 2 or D = D. That is, events are either assigned to a single, a pair, or all daughter processes. For sets D with D = 2, we set 5 QD ∼ N (0, Σ) - a Gaussian distributions of zero mean and some specified covariance. The choice of the precise pairwise shift distributions is not important. Shifts of events attributed to a single process have no effect on the statistics of the multivariate process. (To see this, note that the integrals with respect to t in Eq. (2) below, for example, may be viewed as a marginalization over shifts applied to events in the first process.) It remains to define the jitter distribution for events common to the entire population of daughter processes, i.e. events marked by D. We will show that we can generate cascading activity, and analytically describe the resulting correlation structure. We generate random vectors Y ∼ QD according to the following rule, for each i = 1, . . . , N : 1. Generate independent random variables ϕi ∼ Exp(αi) where αi > 0. 2. Set Yi =(cid:80)i j=1 ϕj. In particular, note that these shift times satisfy YN ≥ . . . ≥ Y2 ≥ Y1 ≥ 0, indicating the chain-like structure of these joint events. From the definition of the model and our general result (Theorem 1) below, we immediately have that κX ij (τ ), the second order cross-cumulant density for the process (i, j), is given by where c2 ij(τ ) = λp{i,j} (cid:90) κX ij (τ ) = c2 ij(τ ) + cN ij (τ ), {i,j} {i,j}(t, t + τ )dt, q cN ij (τ ) = λpD (cid:90) {i,j} D q (t, t + τ )dt (1) (2) define the contributions to the second order cross-cumulant density by the second-order, Gaussian- jittered events and the population-level events, respectively. The functions qD(cid:48) D indicate the densities associated with the distribution QD, projected to the dimensions of D(cid:48). All statistical quantities are precisely defined in the methods. By exploiting the hierarchical construction of the shift times, we can find an expression for the joint density qD, necessary to explicitly evaluate Eq. (1). For a general N -dimensional distribution, f (y1, . . . , yN ) = f (yNy1, . . . , yN−1)f (yN−1y1, . . . , yN−2)··· f (y2y1)f (y1). (3) Since Y1 ∼ Exp(α1), we have f (y1) = exp [−α1y1] Θ(y1), where Θ(y) is the Heaviside step function. Further, as Yi(Y1, . . . , Yi−1) ∼ Yi−1 + Exp(αi) for i ≥ 2, the conditional densities of the yi's take the form f (yiy1, . . . , yi−1) = f (yiyi−1) = αi exp [−αi(yi − yi−1)] Θ(yi − yi−1), i ≥ 2. Substituting this in to the identity Eq. (3), we have (cid:40) α1 exp [−α1y1](cid:81)N qD(y1, . . . , yN ) = 0 i=2 αi exp [−αi(yi − yi−1)] yN ≥ . . . ≥ y2 ≥ y1 ≥ 0 otherwise . (4) Using Theorem 1 (Eq. (33)) we obtain the N th order cross-cumulant density (see the Methods), κX 1···N (τ1, . . . , τN−1) = λpD qD(t, t + τ1, . . . , t + τN−1) = λpD · i=1 αi+1 exp [−αi+1(τi − τi−1)] τi ≥ τi−1 i = 1, . . . , N − 1 otherwise , (5) (cid:90) (cid:40)(cid:81)N−1 0 6 Figure 3: An example of a six dimensional GTaS model exhibiting synfire-like cascading firing patterns. (A) A raster-plot of spiking activity over a 100ms window. Blue spikes indicate either marginal or pairwise events (i.e., events corresponding to markings for sets D ⊂ D with D ≤ 2. Red spikes indicate population-wide events which have shift-times given by cumulative sums of independent exponentials, as described in the text. Arrows indicate the location of the first spike in the cascade. (B) A second-order cross-cumulant κX 13 (black line) of this model is composed of contributions from two sources: correlations due to second-order markings, which have Gaussian shifts (c2 13 -- dashed red line), and correlations due to the the occurrence of population wide events (cN 13 -- dashed blue line). (C) Density plots of the third-order cross-cumulant density for triplets i) (1, 2, 3) and ii) (1, 2, 4) -- the latter is given explicitly in Eq. (6). System parameters are given in the Appendix. where, for notational convenience, we define τ0 = 0. A raster plot of a realization of this model is shown in Fig. 3A. We note that the cross-cumulant densities of arbitrary subcollections of the counting processes X can be obtained by finding the appropriate marginalization of qD via inte- gration of Eq. (4). In the case that common distributions are used to define the shifts, symbolic calculation environments (i.e., Mathematica) can quickly yield explicit formulas for cross-cumulant densities. Mathematica notebooks for Figure 1 available upon request. As a particular example, we consider the cross-cumulant density of the marginal processes X1, X3. Using Eqs. (2, 4), we find 13(τ ) = λpDΘ(τ ) · cN (cid:40) α2α3 α3−α2 α2α3τ exp [−α2τ ] {exp [−α2τ ] − exp [−α3τ ]} α2 (cid:54)= α3 α2 = α3 . An expression for c2 for all i, j. In Fig. 3B, we plot these contributions, as well as the full covariance density. 13(τ ) may be obtained similarly using Eq. (2) and recalling that Q{i,j} ≡ N (0, Σ) Similar calculations at third order yield, as an example, 124(τ1, τ2) = λpD · κX α4−α3 α2α3α4(τ2 − τ1) exp [−α2τ1 − α3(τ2 − τ1)] exp [−α2τ1]{exp [−α3(τ2 − τ1)] − exp [−α4(τ2 − τ1)]} α3 (cid:54)= α4 α3 = α4 , (6) (cid:40) α2α3α4 7 124(τ1, τ2) is supported only on τ2 ≥ τ1 ≥ 0. Plots of the third- where the cross-cumulant density κX order cross-cumulants for triplets (1, 2, 3) and (1, 2, 4) in this model are shown in Fig. 3C. Note that, for the specified parameters, the conditional distribution of Y4 -- the shift applied to the events of X4 in a joint population event -- given Y2 follows a gamma distribution, whereas Y3Y2 follows an exponential distribution, explaining the differences in the shapes of these two cross-cumulant densities. General cross-cumulant densities of at least third order for the cascading model will have a form similar to that given in Eq. (6), and will contain no signature of the correlation of strictly second order events. This highlights a key benefit of cumulants as a measure of dependence: although they agree with central moments up to third order, we know from Eq. (23) below (or Eq. (22) in the general case) that central moments necessarily exhibit a dependence on lower order statistics. On the other hand, cumulants are "pure" and quantify only dependencies which cannot be inferred from lower order statistics [27]. One useful statistic for analyzing population activity through correlations is the population cumulant density [48]. The second order population cumulant density for cell i is defined by (see the Methods) κX i,pop(τ ) = κX ij (τ ). (cid:88) j(cid:54)=i (cid:88) k(cid:54)=i,j This function is linearly related to the spike-triggered average of the population activity conditioned on that of cell i. In Fig. 4 we show three different second-order population-cumulant functions for the cascading GTaS model of Fig. 3A. When the second order population cumulant for a neuron is skewed to the right of τ = 0 (as is κX 1,pop -- blue line), a neuron tends to precede its partners in pairwise spiking events. Similarly, skewness to the left of τ = 0 (κX 6,pop -- orange line) indicates a neuron which tends to trail its partners in such events. A symmetric population indicates a neuron is a follower and a leader. Taken together, these three second order population cumulants hint at the chain structure of the process. Greater understanding of the joint temporal statistics in a multivariate counting process can be obtained by considering higher-order population cumulant densities. We define the third-order population cumulant density for the pair (i, j) to be κX ij,pop(τ1, τ2) = κX ijk(τ1, τ2). The third-order population cumulant density is linearly related to the spike-triggered population activity, conditioned on spikes in cells i and j separated by a delay τ1. In Fig. 4B,C,D, we present three distinct third-order population cumulant densities. Examining κX 12,pop(τ1, τ2) (panel B), we see only contributions in the region τ2 > τ1 > 0, indicating that the pairwise event 1 → 2 often precedes a third spike elsewhere in the population. The population cumulant κX 34,pop(τ1, τ2) has contributions in two sections of the plane (panel C). Contributions in the region τ2 > τ1 > 0 can be understood following the preceding example, while contributions in the region τ2 < 0 < τ1 imply that the firing of other neurons tends to precede the joint firing event 1 → 2. Lastly, contributions to κX 16,pop(τ1, τ2) (panel D) are limited to 0 < τ2 < τ1, indicating an above chance probability of joint firing events of the form 1 → i → 6, where i indicates a distinct neuron within the population. A distinct advantage of the study of population cumulant densities as opposed to individual cross-cumulant functions in practical applications is related to data (i.e., sample size) limitations. In many practical applications, where the temporal structure of a collection of observed point processes is of interest, we often deal with a small, noisy samples. It may therefore be difficult to 8 estimate third- or higher-order cumulants. Population cumulants partially circumvent this issue by pooling [59, 58, 74] (or summing) responses, to amplify existing correlations and average out the noise in measurements. Figure 4: Population cumulants for the synfire-like cascading GTaS process of Fig. 3. See Eq. (25) for the definition of population cumulants. (A) Second order population cumulant densities for processes 1,3 and 6. Greater mass to the right (resp. left) of τ = 0 indicates that a cell tends to lead (resp. follow) in pairwise-correlated events. (B) Third order population cumulant for processes X1, X2 in the cascading GTaS process. Concentration of the mass in different regions of the plane indicates temporal structure of events correlated between X1, X2 relative to the remainder of the population (see the text). (C) Same as (B), but for processes X3, X4. (D) Same as (B), but for processes X1, X6. System parameters are given in the Appendix. We conclude this section by noting that even cascading GTaS examples can be much more general. For instance, we can include more complex shift patterns, overlapping subassemblies within the population, different temporal processions of the cascade, and more. Timing-selective network The responses of single neurons and neuronal networks in experi- mental [10, 69, 53] and theoretical studies [28, 36, 75, 39, 41] can reflect the temporal structure of their inputs. Here, we present a simple example that shows how a network can be selective to fine temporal features of its input, and how the GTaS model can be used to explore such examples. 9 As a general network model, we consider N leaky integrate-and-fire (LIF) neurons with mem- brane potentials Vi obeying = −Vi + dVi dt N(cid:88) j=1 wij(F ∗ zj)(t) + winxi(t), i = 1, . . . , N. (7) When the membrane potential of cell i reaches a threshold V th, an output spike is recorded and the membrane potential is reset to zero, after which evolution of Vi resumes the dynamics in Eq. (7). Here wij is the synaptic weight of the connection from cell j to i, win is the input weight, and we assume time to be measured in units of membrane time constants. The function −1e−(t−τd)/τsynΘ(t − τd) is a delayed, unit-area exponential synaptic kernel with time- F = τsyn constant τsyn and delay τd. When the membrane potential of a cell reaches threshold, V th, a spike is generated and the membrane potential is reset to zero. The output of the ith neuron is where tj i is the time of the jth spike of neuron i. In addition, the input {xi}N i=1 is (cid:88) j (cid:88) j zi(t) = δ(t − tj i ), xi(t) = δ(t − sj i ), i} correspond to those of a GTaS counting process X. Thus, each input where the event times {sj spike results in a jump in the membrane potential of the corresponding LIF neuron of amplitude win. The particular network we consider will have a ring topology (nearest neighbor-only connectivity) -- specifically, for i, j = 1, . . . , N, we let (cid:40) wij = wsyn 0 i − j mod N ≡ 1 or N − 1 otherwise . We further assume that all neurons are excitatory, so that wsyn > 0. A network of LIF neurons with synaptic delay is a minimal model which can exhibit fine-scale discrimination of temporal patterns of inputs without precise tuning [38]. To exhibit this depen- dence we generate inputs from two GTaS processes. The first (the cascading model ) was described in the preceding example. To independently control the mean and variance of relative shifts we replace the sum of exponential shifts with sums of gamma variates. We also consider a model featuring population-level events without shifts (the synchronous model ), where the distribution QD is a δ distribution at zero in all coordinates. The only difference between the two input models is in the temporal structure of joint events. In particular, the rates, and all long timescale spike count cross-cumulants (equivalent to the total "area" under the cross-cumulant density, see the Methods) of order two and higher are identical for the two processes. We focus on the sensitivity of the network to the temporal cumulant structure of its inputs. In Fig. 5A,B, we present two example rasters of the nearest-neighbor LIF network receiving synchronous (left) and cascading (right) input. In the second case, there is an obvious pattern in the outputs, but the firing rate is also increased. This is quantified in Fig. 5C, where we compare the number of output spikes fired by a network receiving synchronous input (horizontal axis) with the same for a network receiving cascading input (vertical axis), over a large number of trials. On 10 average, the cascading input increases the output rate by a factor of 1.5 over the synchronous inputs -- we refer to this quantity as the cascade amplification factor (CAF). Finally, in Fig. 5D, we study how the the cascade amplification factor depends on the parameters that define the timing of spikes for the cascading inputs. First, we study the dependence on the standard deviation σshift of the gamma variates determining the shift distribution. We note that amplification factors above 1.5 hold robustly (i.e., for a range of shift σshift values). The amplification factors decrease with shift variance. In the inset to panel D, we show how the gain depends on the mean of the shift distribution µshift. On an individual trial, the response intensity will depend strongly on the total number of input spikes. Thus, in order to enforce a fair comparison, the mother process and markings used were identical in each trial of every panel of Fig. 5. Figure 5: (A) Example input (left) and output (right) for the nearest neighbor LIF network receiving input with synchronous input. (B) Same as (A), but for cascading input. (C) Scatter plot of the output spike count of the network receiving synchronous (horizontal axis) and cascading input (vertical axis) with µshift = 2, σshift = 0.3. The red line is the diagonal. (D) Average gain (rate in response to cascading input divided by rate in response to synchronous input) as a function of the standard deviation of the gamma variates which compose the shift vectors for population- level events (µshift was fixed at 2). The red dot indicates the value of σshift used in panel C. Inset shows the same gain as panel D, but for varying the mean of the shift distribution (σshift = 0.3). Spike counts in panels C and D were obtained for trials of length T = 100. Other system parameters are given in the Appendix. These observations have simple explanations in terms of the network dynamics and input statis- tics. Neglecting, for a moment, population-level events, the network is configured so that correla- tions in activity decrease with topographic distance. Accordingly, the probability of finding neurons that are simultaneously close to threshold also decreases with distance. Under the synchronous in- put model, a population-level event results in a simultaneous increase of the membrane potentials of all neurons by an amount win, but unless the input is very strong (in which case every, or almost every, neuron will fire regardless of fine-scale input structure), the set of neurons sufficiently close to threshold to "capitalize" on the input and fire will typically be restricted to a topographically 11 adjacent subset. Neurons which do not fire almost immediately will soon have forgotten about this population-level input. As a result, the output does not significantly reflect the chain-like structure of the inputs (Fig. 5A, right). On the other hand, in the case of the cascading input, the temporal structure of the input and the timescale of synapses can operate synergistically. Consider a pair of adjacent neurons in the ring network, called cells 1 and 2, arranged so that cell 2 is downstream from cell 1 in the direction of the population-level chain events. When cell 1 spikes, it is likely that cell 2 will also have an elevated membrane potential. The potential is further elevated by the delayed synaptic input from cell 1. If cell 1 spikes in response to a population-level chain event, then cell 2 imminently receives an input spike as well. If the synaptic filter and time-shift of the input spikes to each cell align, then the firing probability of cell 2 will be large relative to chance. This reasoning can be carried on across the network. Hence synergy between the temporal structure of inputs and network architecture allows the network to selectively respond to the temporal structure of the inputs (Fig. 5B, right). In [46], the effect of higher order correlations on the firing rate gain of an integrate -- and -- fire neuron was studied by driving single cells using sums of SIP or MIP processes with equivalent firing rates (first order cumulants) and pairwise correlations (second order cumulants). In contrast, in the preceding example, the two inputs have equal long time spike count cumulants, and differ only in temporal correlation structure. An increase in firing rate was due to network interactions, and is therefore a population level effect. We return to this comparison in the Discussion. These examples demonstrate how the GTaS model can be used to explore the impact of spatio- temporal structure in population activity on network dynamics. We next proceed with a formal derivation of the cumulant structure for a general GTaS process. 2.3 Cumulant structure of a GTaS process The GTaS model defines an N -dimensional counting process. Following the standard description for a counting process, X = (X1, . . . , XN ) on RN , given a collection of Borel subsets Ai ∈ B(R), i = 1, . . . , N , then X(A1 × ··· × AN ) = (X1(A1), . . . , XN (AN )) ∈ NN is a random vector where the value of each coordinate i indicates the (random) number of points which fall inside the set Ai. Note that the GTaS model defines processes that are marginally Poisson. For each D ⊂ D = {1, . . . , N}, define the tail probability ¯pD by (cid:88) D⊂D(cid:48)⊂D ¯pD = pD(cid:48). (8) Since pD is the probability that exactly the processes in D are marked, ¯pD is the probability that all processes in D, as well as possibly other processes, are marked. An event from the mother process is assigned to daughter process Xi with probability ¯p{i}. As noted above, an event attributed to process i following a marking D (cid:51) i will be marginally shifted by a random amount determined {i} by the distribution Q D which represents the projection of QD onto dimension i. Thus, the events in the marginal process Xi are shifted in an independent and identically distributed (IID) manner according to the mixture distribution Qi given by (cid:80) (cid:80) {i} D(cid:51)i pDQ D D(cid:51)i pD . Qi = Note that IID shifting of the event times of a Poisson process generates another Poisson process of identical rate. Thus, the process Xi is marginally Poisson with rate λ¯p{i} [60]. 12  .  In deriving the statistics of the GTaS counting process X, it will be useful to express the distribution of X as  X1(A1) ...  =distr XN (AN )  (cid:80) (cid:80) D(cid:51)1 ξ(D; A1, . . . , AN ) ... D(cid:51)N ξ(D; A1, . . . , AN ) Here, each ξ(D; A1, . . . , AN ) is an independent Poisson process. This process counts the number of points which are marked by a set D(cid:48) ⊃ D, but (after shifting) only the points with indices i ∈ D lie in the corresponding set Ai. Precise definitions of the processes ξ and a proof of Eq. (9) may be found in the Appendix. We emphasize that the Poisson processes ξ(D) do not directly count points marked for the set D, but instead points which are marked for a set containing D that, after shifting, only have their D-components lying in the "relevant" sets Ai. {Xij}ij∈ ¯D for some set ¯D ⊂ D, consisting of ¯D = k distinct members of the collection of counting Suppose we are interested in calculating dependencies among a subset of daughter processes, processes X. Then the following alternative representation will be useful: Xi1(Ai1) ...  =distr  (cid:80) (cid:80) (cid:88) D(cid:48)⊃D ( ¯D\D)∩D(cid:48)=∅ i1∈D⊂ ¯D ζD(A1, . . . , AN ) ... ik∈D⊂ ¯D ζD(A1, . . . , AN ) ξ(D(cid:48); A1, . . . , AN ). where Xik (Aik ) ζD(A1, . . . , AN ) = We illustrate this decomposition in the cases k = 2, 3 in Fig. 6. The sums in Eq. (10) run over all sets D ⊂ D containing the indicated indices ij and contained within ¯D. The processes ζD are comprised of a sum of all of the processes ξ(D(cid:48)) (defined below Eq. (9)) such that D(cid:48) contains all of the indices D, but no other indices which are part of the subset ¯D under consideration. These sums are non-overlapping, implying that the ζD are also independent and Poisson. The following examples elucidate the meaning and significance of Eq. (10). We emphasize that the GTaS process is a completely characterized, joint Poisson process, and we use Eq. (10) to calculate cumulants of a GTaS process. In principle, any other statistics can be obtained similarly. (9) (10) Second order cumulants (covariance) We first generalize a well-known result about the de- pendence structure of temporally jittered pairs of Poisson processes, X1, X2. Assume that events from a mother process with rate λ, are assigned to two daughter processes with probability p. Each event time is subsequently shifted independently according to a univariate distribution f . The cross-cumulant density (or cross-covariance function; see the Methods for cumulant definitions) then has the form [14] We generalize this result within the GTaS framework. At second order, Eq. (10) has a particularly nice form. Following [11] we write for i (cid:54)= j (see Fig. 6A) . (11) (cid:90) κX 12(τ ) = λp f (t)f (t + τ )dt = λp(f (cid:63) f )(τ ). (cid:18)Xi(Ai) (cid:19) Xj(Aj) =distr (cid:18)ζ{i,j}(Ai, Aj) + ζ{i}(Ai) (cid:19) ζ{i,j}(Ai, Aj) + ζ{j}(Aj) 13 Figure 6: (A) Illustrating the representation given by Eq. (10) in the case of two distinct processes (see Eq. (11)) with N = 4 and ¯D = {1, 2}. (B) Same as (A), for three processes with ¯D = {1, 2, 3} (see Eq. (16)). The process ζ{i,j} sums all ξ(D(cid:48)) for which {1, 2} ⊂ D(cid:48), while the process ζ{i} sums all ξ(D(cid:48)) such that i ∈ D(cid:48), j /∈ D(cid:48), and ζ{j} is defined likewise. Using the representation in Eq. (11), we can derive the second order cumulant (covariance) structure of a GTaS process. First, we have cov[Xi(Ai), Xj(Aj)] = κ[Xi(Ai), Xj(Aj)] = κ[ζ{i,j}(Ai, Aj), ζ{i,j}(Ai, Aj)] + κ[ζ{i}(Ai), ζ{i,j}(Ai, Aj)] + κ[ζ{i,j}(Ai, Aj), ζ{j}(Aj)] + κ[ζ{i}(Ai), ζ{j}(Aj)] = κ2[ζ{i,j}(Ai, Aj)] + 0 = E(cid:2)ζ{i,j}(Ai, Aj)(cid:3) . The third equality follows from the construction of the processes ζD: if D (cid:54)= D(cid:48), then the processes ζD, ζD(cid:48) are independent. The final equality follows from the observation that every cumulant of a Poisson random variable equals its mean. The covariance may be further expressed in terms of model parameters (see Theorem 1 for a (cid:90) (cid:88) D(cid:48)⊃{i,j} generalization of this result to arbitrary cumulant orders): cov[Xi(Ai), Xj(Aj)] = λ pD(cid:48) P (t + Yi ∈ Ai, t + Yj ∈ Aj Y ∼ QD(cid:48)) dt. (12) In other words, the covariance of the counting processes is given by the weighted sum of the probabilities that the (i, j) marginal of the shift distributions yield values in the appropriate sets. 14 The weights are the intensities of each corresponding component processes ξ(D) which contribute events to both of the processes i and j. In the case that QD ≡ Q, Eq. (12) reduces to the solution given in [11]. Using the tail probabilities defined in Eq. (8), if QD ≡ Q for all D, the integral in Eq. (12) no longer depends on the subset D(cid:48), and the equation may be written as cov[Xi(Ai), Xj(Aj)] = λ¯p{i,j} P (t + Yi ∈ Ai, t + Yj ∈ Aj Y ∼ Q) dt. Using Eq. (12), we may also compute the second cross-cumulant density (also called the covari- ance density) of the processes. From the definition of the cross-cumulant density (Eq. (24) in the Methods), this is given by cov[Xi([0, ∆t)), Xj([τ, τ + ∆t))] κX ij (τ ) = lim ∆t→0 (cid:88) (cid:90) = λ D(cid:48)⊃{i,j} pD(cid:48) lim ∆t→0 ∆t2 P (t + Yi ∈ [0, ∆t), t + Yj ∈ [τ, τ + ∆t) Y ∼ QD(cid:48)) (13) dt. ∆t2 Before continuing, we note that given a random vector Y = (Y1, . . . , YN ) ∼ Q, where Q has density q(y1, . . . , yN ), the vector Z = (Y2 − Y1, . . . , YN − Y1) has density qZ given by (cid:90) (cid:90) qZ(τ1, . . . , τN−1) = q(t, t + τ1, . . . , t + τN−1)dt. (14) Assuming that the distributions QD(cid:48) have densities qD(cid:48), and denoting by q density of the variables Yi, Yj under QD(cid:48), we have that {i,j} D(cid:48) the bivariate marginal κX ij (τ ) = λ pD(cid:48) {i,j} D(cid:48) q (t, t + τ )dt. (15) (cid:90) (cid:88) D(cid:48)⊃{i,j} According to Eq. (14), the integrals present in Eq. (15) are simply the densities of the variables Yj − Yi, where Y ∼ QD(cid:48). Thus κX ij (τ ), which captures the additional probability for events in the marginal processes Xi and Xj separated by τ units of time beyond what can be predicted from lower order statistics is given by a weighted sum (in this case, the lower order statistics are marginal intensities -- see the discussion around Eq. (24) of the Methods). The weights are the "marking rates" λpD(cid:48) for markings contributing events to both component processes, while the summands are the probabilities that the corresponding shift distributions yield a pair of shifts in the proper arrangement - specifically, the shift applied to the event as attributed to Xi precedes that applied to the event mapped to Xj by τ units of time. Third order cumulants To determine the higher order cumulants for a GTaS process, one can again use the representation given in Eq. (10). The distribution of a subset of three processes may be expressed in the form (see Fig. 6B)  Xi(Ai) Xj(Aj) Xk(Ak)  =distr  ζ{i,j,k} + ζ{i,j} + ζ{i,k} + ζ{i} ζ{i,j,k} + ζ{i,j} + ζ{j,k} + ζ{j} ζ{i,j,k} + ζ{i,k} + ζ{j,k} + ζ{k},  , (16) 15 where, for simplicity, we suppressed the arguments of the different ζD on the right hand side. Again, the processes in the representation are independent and Poisson distributed. The variable ζ{i,j,k} is the sum of all random variables ξ(D) (see Eq. (9)) with D ⊃ {i, j, k}, while the variable ζ{i,j} is now the sum of all ξ(D) with D ⊃ {i, j}, but k /∈ D. The rest of the variables are defined likewise. Using properties (C1) and (C2) of cumulants given in the Methods, and assuming that i, j, k are distinct indices, we have κ(Xi(Ai), Xj(Aj), Xk(Ak)) = κ3(ζ{i,j,k}) = E(cid:2)ζ{i,j,k}(cid:3) . The second equality follows from the fact that all cumulants of a Poisson distributed random variable equal its mean. Similar to Eq. (12), we may write κ(Xi(Ai), Xj(Aj), Xk(Ak)) = λ pD(cid:48) P (t + Yi ∈ Ai, t + Yj ∈ Aj, t + Yk ∈ Ak Y ∼ QD(cid:48)) dt. The third cross-cumulant density is then given similarly to the second order function by κX ijk(τ1, τ2) = λ pD(cid:48) {i,j,k} D(cid:48) q (t, t + τ1, t + τ2)dt. (cid:90) (cid:88) D(cid:48)⊃{i,j,k} (cid:90) (cid:88) D(cid:48)⊃{i,j,k} Here, we have again assumed the existence of densities qD(cid:48), and denote by q the joint marginal density of the variables Yi, Yj, Yk under qD(cid:48). The integrals appearing in the expression for the third order cross-cumulant density are the probability densities of the vectors (Yj − Yi, Yk − Yi), where Y ∼ QD(cid:48). {i,j,k} D(cid:48) General cumulants Finally, consider a general subset of k distinct members of the vector count- ing process X as in Eq. (10). The following theorem provides expressions for the cross-cumulants of the counting processes, as well as the cross-cumulant densities, in terms of model parameters in this general case. The proof of Theorem 1 is given in the Appendix. Theorem 1. Let X be a joint counting process of GTaS type with total intensity λ, marking distribution (pD)D⊂D, and family of shift distributions (QD)D⊂D. Let A1, . . . , Ak be arbitrary sets in B(R), and ¯D = {i1, . . . , ik} ⊂ D with ¯D = k. The cross-cumulant of the counting processes may be written κ(Xi1(A1), . . . , Xik (Ak)) = λ pD(cid:48) P (t1 + Y ¯D ∈ A1 × ··· × AkY ∼ QD(cid:48))dt (17) where Y ¯D represents the projection of the random vector Y on to the dimensions indicated by the members of the set ¯D. Furthermore, assuming that the shift distributions possess densities (qD)D⊂D, the cross-cumulant density is given by κX i1···ik (τ1, . . . , τk−1) = λ pD(cid:48) ¯D D(cid:48)(t, t + τ1,··· , t + τk−1)dt, q (18) (cid:90) (cid:88) D(cid:48)⊃ ¯D (cid:90) (cid:88) D(cid:48)⊃ ¯D where q ¯D D(cid:48) indicates the kth order joint marginal density of qD(cid:48) in the dimensions of ¯D. 16 An immediate corollary of Theorem 1 is a simple expression for the infinite-time-window cu- mulants, obtained by integrating the cumulant density across all time lags τi. From Eq. (33), we have (cid:90) (cid:90) (cid:88) D(cid:48)⊃ ¯D γX i1···ik (∞) = ··· κX i1···ik (τ1, . . . , τk−1)dτk−1 ··· dτ1 = λ pD(cid:48) · 1 = λ¯p ¯D. (19) This shows that the infinite time window cumulants for a GTaS process are non-increasing with respect to the ordering of sets, i.e., γX i1···ik (∞) ≥ γX i1···ikik+1 (∞). We conclude this section with a short technical remark: Until this point, we have considered only the cumulant structure of sets of unique processes. However occasionally, one may wish to calculate a cumulant for a set of processes including repeats. Take, for example, a cumulant κ(X1(A1), X1(A2), X3(A3)). Owing to the marginally Poisson nature of the GTaS process, we would have (referring to the Methods for cumulant definitions) κ(X1(A1), X1(A2), X3(A3)) = κ(2,1)(X1(A1 ∩ A2), X3(A3)) if X ∼ GTaS. For a general counting process X, it may be shown that κX 113(τ1, τ2) = δ(τ1)κX 13(τ2) + "non-singular contributions". In addition, the second order auto-cumulant density may be written [21] κX ii (τ ) = riδ(τ ) + "non-singular contributions", (20) (21) where ri is the stationary rate. The singular contribution shown in Eq. (21) at third order is in analogy to the delta contribution proportional to the firing rate which appears in the second- order auto-cumulant density. For a GTaS process, the non-singular contributions in Eq. (21) are identically zero, following directly from Eq. (20). Expressions similar to Eqs. (20, 21) hold for general cases. 3 Discussion We have introduced a general method of generating spike trains with flexible spatiotemporal struc- ture. The GTaS model is completely analytically tractable: all statistics of interest can be obtained directly from the distributions used to define it. It is based on an intuitive method of selecting and shifting point processes from a "mother" train. Moreover, the GTaS model can be used to easily generate partially synchronous states, cluster firing, cascading chains, and other spatiotemporal patterns of neural activity. Processes generated by the GTaS model are naturally described by cumulant densities of pair- wise and higher orders. This raises the question of whether such statistics are readily computable from data, so that realistic classes of GTaS models can be defined in the first place. One approach is to fit mechanistic models to data, and to use the higher order structure that follows from the underlying mechanisms [79]. A synergistic blend of other methods with the GTaS framework may also be fruitful -- for example, the CuBIC framework of [70] could be used to determine relevant marking orders, and the parametrically-described GTaS process could then be fit to allow gen- eration of surrogate data after selection of appropriate classes of shift distributions. When it is 17 necessary to infer higher order structure in the face of data limitations, population cumulants are an option to increase statistical power (albeit at the cost of spatial resolution; see Figure 4). While the GTaS model has flexible higher order structure, it is always marginally Poisson. While throughout the cortex, spiking is significantly irregular [65, 35], the level of variability differs across cells, with Fano factors ranging from below 0.5 to above 1.5 -- in comparison with the Poisson value of 1 [20]. Changes in variability may reflect cortical states and computation [77, 47]. A model that would allow flexible marginal variability would therefore be very useful. Unfortunately, the tractability of the GTaS model is closely related to the fact that the marginal processes are Poisson. Therefore an immediate generalization does not seem possible. A number of other models have been used to describe population activity. Maximum entropy (ME) approaches also result in models with varied spatial activity; these are defined based on moments or other averaged features multivariate spiking activity [63, 61]. Such models are often used to fit purely spatial patterns of activity, though [72, 52] have extended the techniques to treat temporal correlations as well. Generalized linear models (GLMs) have been used successfully to describe spatiotemporal patterns at second [57], and third order [56]. In comparison to the present GTaS method, both GLMs and ME models are more flexible. They are feature well-defined approaches for fitting to data, including likelihood-based methods with well-behaved convexity properties. What the GTaS method contributes is an explicit way to generate population activity with explicitly specified high order spatio-temporal structure. Moreover, the lower order cumulant structure of a GTaS process can be modified independently of the higher order structure, though the reverse is not true. There are a number of possible implications of such spatio-temporal structure for communication within neural networks. In Section 2.2, we showed that these temporal correlations can play a role similar to that of spatial correlations established in [46] for determining network input-output transfer. Our model allowed us to examine that impact of such temporal correlations on the network-level gain of a downstream population (cascade amplification factor). Even in a very simple network it was clear that the strength of the response is determined jointly by the temporal structure of the input to the network, and the connectivity within the network. Kuhn et al. examined the effect of higher order structure on the firing rate gain of an integrate -- and -- fire neuron by driving it with a mixture of SIP or MIP processes [46]. However, in these studies, only the spatial structure of higher order activity was varied. The GTaS model allows us to concurrently change the temporal structure of correlations. In addition, the precise control of the cumulants allows us to derive models which are equivalent up to a certain cross-cumulant order, when the configuration of marking probabilities and shift distributions allow it (as for the SIP and MIP processes of [46], which are equivalent at second order). Such patterns of activity may be useful when experimentally probing dendritic information processing [26], or investigating the response of neuronal networks to complex patterns of input [42]. Spatiotemporal patterns may also be generated by cell assemblies [10]. The firing in such assemblies can be spatially structured, and this structure may not be reflected in the activity of participating cells. Assemblies can exhibit persistent patterns of firing, sometimes with millisecond precision [33]. The GTaS framework is well suited to describe exactly such activity patterns. The examples we presented can be easily extended to generate more complex patterns of activity with overlapping cell assemblies, different cells leading the activity, and other variations. Understanding impact of spatiotemporal patterns on neural computations remains an open and exciting problem. Progress will require cooperation among simulation, theory, and experimental work -- the latter taking advantage of novel stimulation techniques. We hope that the GTaS model, 18 as a practical and flexible method for generating high-dimensional, correlated spike trains, will play a significant role along the way. 4 Methods Cumulants as a measure of dependence We first define cross-cumulants (also called joint cumulants) [71, 25, 43] and review some important properties of these quantities. Define the cumulant generating function g of a random vector X = (X1, . . . , XN ) by g(t1, . . . , tN ) = log  N(cid:88) E exp  . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)t1=···=tN =0 where r = (r1, . . . , rN ) is a N -vector of positive integers, and r =(cid:80)N The r-cross-cumulant of the vector X is given by ∂r 1 ··· ∂trN ∂tr1 N g(t1, . . . , tN ) κr(X) = . tjXj j=1 i=1 ri. We will generally deal with cumulants where all variables are considered at first order, without excluding the possibility that some variables are duplicated. In this case, we define the cross-cumulant κ(X), of the variables in the random vector X = (X1, . . . , XN ) as κ(X) := κ1(X) = ∂N ∂t1 ··· ∂tN This relationship may be expressed in combinatorial form: g(t1, . . . , tN ) (cid:12)(cid:12)(cid:12)(cid:12)t1=···=tN =0 (cid:88) (π − 1)!(−1)π−1 (cid:89) where 1 = (1, . . . , 1). (cid:35) Xi (22) (cid:34)(cid:89) i∈B E B∈π κ(X1, . . . , XN ) = π (cid:88) where π runs through all partitions of D = {1, . . . , N}, and B runs over all blocks in a partition π. More generally, the r-cross-cumulant may be expressed in terms of moments by expanding the cumulant generating function as a Taylor series, noting that g(t1, . . . , tN ) = κr(X1, . . . , XN ) 1 ··· xrN xr1 d with r! = ri!, r r! N(cid:89) i=1 similarly expanding the moment generating function M (t) = eg(t), and matching the polynomial coefficients. Note that the nth cumulant κn of a random variable X may be expressed as a joint cumulant via κn(X) = κ(X, . . . , X) . (cid:124) (cid:123)(cid:122) (cid:125) n copies of X We will utilize the following two principal properties of cumulants [71, 15, 70, 54]: (C1) Multilinearity - for any random variables X, Y,{Zi}N i=2, we have κ(aX + bY, Z2, . . . , ZN ) = aκ(X, Z2, . . . , ZN ) + bκ(Y, Z2, . . . , ZN ). This holds regardless of dependencies amongst the random variables. 19 (C2) If any subset of the random variables in the cumulant argument is independent from the remaining, the cross-cumulant is zero - i.e., if {X1, . . . , XN1} and {Y1, . . . , YN2} are sets of random variables such that each Xi is independent from each Yj, then for all rX ∈ NN1 κ(rX ,rY )(X1, . . . , XN1, Y1, . . . , YN2) = 0 + , rY ∈ NN2 + . To exhibit another key property of cumulants, consider a 4-vector X = (X1, X2, X3, X4) with non-zero fourth cumulant and a random variable Z independent of each Xi. Define Y = (X1 + Z, X2 + Z, X3 + Z, X4). Using properties (C1), (C2) above, it follows that κ(Y1, Y2, Y3) = κ(X1, X2, X3) + κ3(Z). On the other hand, it is also true that κ(Y) = κ(X), that is, adding the variable Z to only a subset of the variables in X results in changes to cumulants involving only that subset, but not to the joint cumulant of the entire vector. In this sense, an rth order cross-cumulant of a collection of random variables captures exclusively dependencies amongst the collection which cannot be described by cumulants of lower order. In the example above, only the joint statistical properties of a subset of X were changed. As a result, the total cumulant κ(X) remained fixed. From Eq. (22), it is apparent that κ(Xi) = E[Xi], and κ(Xi, Xj) = cov[Xi, Xj]. In addition, the third cumulant, like the second, is equal to the corresponding central moment: κ(Xi, Xj, Xk) = E[(Xi − E[Xi])(Xj − E[Xj])(Xk − E[Xk])] . As cumulants and central moments agree up to third order, central moments up to third order inherit the properties discussed above at these orders. On the other hand, the fourth cumulant is not equal to the fourth central moment. Rather: κ(Xi,Xj, Xk, Xl) = E[(Xi − E[Xi])(Xj − E[Xj])(Xk − E[Xk])(Xl − E[Xl])] − cov[Xi, Xj] cov[Xk, Xl] − cov[Xi, Xk] cov[Xj, Xl] − cov[Xi, Xl] cov[Xj, Xk] . (23) Higher cumulants have similar (but more complicated) expansions in terms of central moments. Accordingly, central moments of fourth and higher order do not inherit properties (C1), (C2). Temporal statistics of point processes In the Results, we present an extension of previous work [11] in which we construct and analyze multivariate counting processes X = (X1, . . . , XN ) where each Xi is marginally Poisson. Formally, a counting process X is an integer-valued random measure on B(RN ). Evaluated on subset A1 × ··· × AN of B(RN ), the random vector (X1(A1), . . . , XN (AN )) counts events in d distinct categories whose times of occurrence fall in to the sets Ai. A good general reference on the properties of counting processes (marginally Poisson and otherwise) is [22]. The assumption of Poisson marginals implies that for a set Ai ∈ B(R), the random variable Xi(Ai) follows a Poisson distribution with mean λi(cid:96)(Ai), where (cid:96) is the Lebesgue measure on R, and λi is the (constant) rate for the ith process. The processes under consideration will further satisfy a joint stationarity condition, namely that the distribution of the vector (X1(A1 +t), . . . , XN (AN +t)) does not depend on t, where Ai + t denotes the translated set {a + t : a ∈ Ai}. 20 We now consider some common measures of temporal dependence for jointly stationary vector counting processes. We will refer to the quantity Xi[0, T ] as the spike count of process i over [0, T ]. The quantity γX (T ) (which we will refer to as a spike count cumulant) is given by i1···ik γX i1···ik (T ) = 1 T κ[Xi1[0, T ], . . . , Xik [0, T ]] measures kth order correlations amongst spike counts for the listed processes which occur over windows of length T . At second order, γX ij (T ) measures the covariance of the spike counts of processes i, j over a common window of length T . The infinite window spike count cumulant quantifies dependencies in the spike counts of point processes over arbitrarily long windows, and is given by γX i1···ik (∞) = lim T→∞ γX i1···ik (T ). A related measure is the kth order cross-cumulant density κX i1,...,ik (τ1, . . . , τk−1), defined by κX i1···ik (τ1, . . . , τk−1) = lim ∆t→0 1 ∆tk κ[Xi1[0, ∆t], Xi2[τ1, τ1 + ∆t], . . . , Xik [τk−1, τk−1 + ∆t]]. (24) The cross-cumulant density should be interpreted as a measure of the likelihood -- above what may be expected from knowledge of the lower order cumulant structure -- of seeing events in processes i2, . . . , ik at times τ1 + t, . . . , τk−1 + t, conditioned on event in process i1 at time t. The infinite window spike count cumulant is equal to the total integral under the cross-cumulant density, (cid:90) (cid:90) γX i1···ik (∞) = ··· κX i1···ik (τ1, . . . , τk−1)dτk−1 ··· dτ1. As an example, we again consider the familiar second-order cross-cumulant density κX ij (τ ) - often referred to as the cross-covariance density or cross-correlation function. Defining the conditional intensity hij(τ ) of process j, conditioned on process i to be hX ij (τ ) = lim ∆t→0 1 ∆t P (Xj[τ, τ + ∆t] > 0Xi[0, ∆t] > 0), that is, the intensity of j conditioned on an event in process i which occurred τ units of time in the past, then it is not difficult to show that ij (τ ) = λihij(τ ) − λiλj. κX That is, the second order cross-cumulant density supplies the probability of chance of observing an event attributed to process i, followed by one attributed to process j, τ units of time later, above what would be expected from knowledge of first order statistics (given by the product of the marginal intensities, λiλj). More generally, at higher orders, the cross-cumulant density should be interpreted as a measure of the likelihood (above what may be expected from knowledge of the lower order correlation structure) of seeing events attribute to processes i2, . . . , ik at times τ1 + t, . . . , τk−1 + t, conditioned on an event in process i1 at time t. Another statistic useful in the study of a correlated vector counting process X is the population cumulant density. At second-order, the population cumulant density for Xi takes the form [48] κX i,pop(τ ) = κX ij (τ ). More generally, the kth order population cumulant density corresponding to the processes Xi1, . . . , Xik−1 is given by κX i1···ik−1,pop(τ1, . . . , τk−1) = κX i1···ik−1j(τ1, . . . , τk−1). (25) (cid:88) (cid:88) j(cid:54)=i j(cid:54)=i1,...,ik 21 References [1] M Abeles and Y Prut. Spatio-temporal firing patterns in the frontal cortex of behaving monkeys. J Physiology-Paris, 90(3):249 -- 250, 1996. [2] Moshe Abeles. Corticonics: Neural circuits of the cerebral cortex. Cambridge University Press, 1991. [3] AMHJ Aertsen, M Diesmann, and MO Gewaltig. Propagation of synchronous spiking activity in feedforward neural networks. J Physiology-Paris, 90(3):243 -- 247, 1996. [4] Shun-ichi Amari, Hiroyuki Nakahara, Si Wu, and Yutaka Sakai. Synchronous firing and higher- order interactions in neuron pool. Neural Comput, 15(1):127 -- 142, 2003. [5] AM Amjad, DM Halliday, JR Rosenberg, and BA Conway. An extended difference of coherence test for comparing and combining several independent coherence estimates: theory and appli- cation to the study of motor units and physiological tremor. J Neurosci Meth, 73(1):69 -- 79, 1997. [6] Bruno B Averbeck, Peter E Latham, and Alexandre Pouget. Neural correlations, population coding and computation. Nat Rev Neurosci, 7(5):358 -- 366, May 2006. [7] Yuval Aviel, Evgeny Pavlov, Moshe Abeles, and David Horn. Synfire chain in a balanced network. Neurocomputing, 44:285 -- 292, 2002. [8] W. Bair, E. Zohary, and W.T. Newsome. Correlated firing in macaque visual area mt: time scales and relationship to behavior. J Neurosci, 21(5):1676 -- 1697, 2001. [9] Andrea K Barreiro, Julijana Gjorgjieva, Fred Rieke, and Eric Shea-Brown. When are feedfor- ward microcircuits well-modeled by maximum entropy methods? Arxiv preprint, 2010. [10] Brice Bathellier, Lyubov Ushakova, and Simon Rumpel. Discrete neocortical dynamics predict behavioral categorization of sounds. Neuron, 76(2):435 -- 449, 2012. [11] N. Bauerle and R. Grubel. Multivariate counting processes: copulas and beyond. Astin Bulletin, 35(2):379, 2005. [12] Tiago Branco, Beverley A Clark, and Michael Hausser. Dendritic discrimination of temporal input sequences in cortical neurons. Sci Signal, 329(5999):1671, 2010. [13] Tiago Branco and Michael Hausser. Synaptic integration gradients in single cortical pyramidal cell dendrites. Neuron, 69(5):885 -- 892, 2011. [14] Romain Brette. Generation of correlated spike trains. Neural Comput, 21(1):188 -- 215, 2009. [15] David R Brillinger. An introduction to polyspectra, 1964. [16] Gyorgy Buzs´aki. Neural syntax: cell assemblies, synapsembles, and readers. Neuron, 68(3):362 -- 385, 2010. [17] Nicholas Cain and Eric Shea-Brown. Impact of correlated neural activity on decision-making performance. Neural Comput, 25(2):289 -- 327, 2013. 22 [18] Catherine E Carr, Hagai Agmon-Snir, and John Rinzel. The role of dendrites in auditory coincidence detection. Nature, 393(6682):268 -- 272, May 1998. [19] Brian Y Chow, Xue Han, Allison S Dobry, Xiaofeng Qian, Amy S Chuong, Mingjie Li, Michael A Henninger, Gabriel M Belfort, Yingxi Lin, Patrick E Monahan, et al. High- performance genetically targetable optical neural silencing by light-driven proton pumps. Na- ture, 463(7277):98 -- 102, 2010. [20] Mark M Churchland, Byron M Yu, John P Cunningham, Leo P Sugrue, Marlene R Cohen, Greg S Corrado, William T Newsome, Andrew M Clark, Paymon Hosseini, Benjamin B Scott, David C Bradley, Matthew A Smith, Adam Kohn, J Anthony Movshon, Katherine M Arm- strong, Tirin Moore, Steve W Chang, Lawrence H Snyder, Stephen G Lisberger, Nicholas J Priebe, Ian M Finn, David Ferster, Stephen I Ryu, Gopal Santhanam, Maneesh Sahani, and Krishna V Shenoy. Stimulus onset quenches neural variability: a widespread cortical phe- nomenon. Nat Neurosci, 13(3):369 -- 378, March 2010. [21] D David Roxbee Cox and Valerie Isham. Point processes, volume 12. Chapman & Hall/CRC, 1980. [22] DJ Daley and D Vere-Jones. An Introduction to the Theory of Point Processes: Volume I: Elementary Theory and Methods, volume 1. Springer, 2002. [23] Gaddum Duemani Reddy, Keith Kelleher, Rudy Fink, and Peter Saggau. Three-dimensional random access multiphoton microscopy for functional imaging of neuronal activity. Nat Neu- rosci, 11(6):713 -- 720, June 2008. [24] Elad Ganmor, Ronen Segev, and Elad Schneidman. Sparse low-order interaction network underlies a highly correlated and learnable neural population code. Proc Natl Acad Sci, 108(23):9679 -- 9684, 2011. [25] CW Gardiner. Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sci- ences. Springer-Verlag, Berlin, 2009. [26] Sonia Gasparini and Jeffrey C Magee. State-dependent dendritic computation in hippocampal CA1 pyramidal neurons. J Neurosci, 26(7):2088 -- 2100, February 2006. [27] Sonja Grun and Stefan Rotter. Analysis of parallel spike trains. Springer, 2010. [28] Robert Gutig and Haim Sompolinsky. The tempotron: a neuron that learns spike timing -- based decisions. Nat Neurosci, 9(3):420 -- 428, 2006. [29] Diego A Gutnisky and Kresimir Josic. Generation of spatio-temporally correlated spike-trains and local-field potentials using a multivariate autoregressive process. J Neurophysiol, December 2009. [30] Xue Han and Edward S Boyden. Multiple-color optical activation, silencing, and desynchro- nization of neural activity, with single-spike temporal resolution. PloS one, 2(3):e299, 2007. [31] Bryan J Hansen, Mircea I Chelaru, and Valentin Dragoi. Correlated variability in laminar cortical circuits. Neuron, 76(3):590 -- 602, 2012. [32] Kenneth D Harris. Neural signatures of cell assembly organization. Nat Rev Neurosci, 6(5):399 -- 407, 2005. 23 [33] Kenneth D Harris, Darrell A Henze, Hajime Hirase, Xavier Leinekugel, George Dragoi, Andras Czurk´o, Gyorgy Buzs´aki, et al. Spike train dynamics predicts theta-related phase precession in hippocampal pyramidal cells. Nature, 417(6890):738 -- 741, 2002. [34] Donald Olding Hebb. The organization of behavior: A neuropsychological theory. Psychology Press, 1949. [35] G. R. Holt, W. R. Softky, C. Koch, and R. J. Douglas. Comparison of discharge variability in vitro and in vivo in cat visual cortex neurons. J Neurophysiol, 75(5):1806 -- 1814, 1996. [36] JJ Hopfield. Pattern recognition computation using action potential timing for stimulus rep- resentation. Nature, 376(6535):33 -- 36, 1995. [37] Yuji Ikegaya, Gloster Aaron, Rosa Cossart, Dmitriy Aronov, Ilan Lampl, David Ferster, and Rafael Yuste. Synfire chains and cortical songs: temporal modules of cortical activity. Science Sig, 304(5670):559, 2004. [38] Eugene M Izhikevich. Polychronization: Computation with spikes. Neural Comput, 18(2):245 -- 282, 2006. [39] Lloyd A Jeffress. A place theory of sound localization. J Comp Physiol Psychol, 41(1):35 -- 39, 1948. [40] Don H Johnson and Ilan N Goodman. Jointly poisson processes. arXiv preprint, 2009. [41] P X Joris, P H Smith, and TCT Yin. Coincidence detection in the auditory system: 50 years after Jeffress. Neuron, 21(6):1235 -- 1238, December 1998. [42] I Kahn, U Knoblich, M Desai, J Bernstein, AM Graybiel, ES Boyden, RL Buckner, and CI Moore. Optogenetic drive of neocortical pyramidal neurons generates fMRI signals that are correlated with spiking activity. Brain Res, 2013. [43] Maurice G Kendall, Alan Stuart, and JK Ord. The advanced theory of statistics (volume 1), 1969. [44] U. Koster, J. Sohl-Dickstein, C. Gray, and B. Olshausen. Higher Order Correlations within arXiv preprint q- Cortical Layers Dominate Functional Connectivity in Microcolumns. bio/1301.0050, 2013. [45] Michael Krumin and Shy Shoham. Generation of spike trains with controlled auto- and cross- correlation functions. Neural Comput, 21(6):1642 -- 1664, June 2009. [46] A. Kuhn, A. Aertsen, and S. Rotter. Higher-order statistics of input ensembles and the response of simple model neurons. Neural Comput, 15(1):67 -- 101, 2003. [47] Ashok Litwin-Kumar and Brent Doiron. Slow dynamics and high variability in balanced cortical networks with clustered connections. Nat Neurosci, 15(11):1498 -- 1505, September 2012. [48] Artur Luczak, Peter Bartho, and Kenneth D Harris. Gating of sensory input by spontaneous cortical activity. J Neurosci, 33(4):1684 -- 1695, January 2013. 24 [49] Artur Luczak, Peter Barth´o, Stephan L Marguet, Gyorgy Buzs´aki, and Kenneth D Harris. Se- quential structure of neocortical spontaneous activity in vivo. Proc Natl Acad Sci, 104(1):347 -- 352, 2007. [50] Jakob H Macke, Philipp Berens, Alexander S Ecker, Andreas S Tolias, and Matthias Bethge. Generating spike trains with specified correlation coefficients. Neural Comput, 21(2):397 -- 423, February 2009. [51] Jakob H Macke, Manfred Opper, and Matthias Bethge. Common input explains higher-order correlations and entropy in a simple model of neural population activity. Phys Rev Lett, 106(20):208102, 2011. [52] O. Marre, S. El Boustani, Y. Fre?gnac, and A. Destexhe. Prediction of spatiotemporal patterns of neural activity from pairwise correlations. Physical Review Letters, 102:138101, 2009. [53] M Meister and MJ Berry II. The neural code of the retina. Neuron, 22(3):435, 1999. [54] J.M. Mendel. Tutorial on higher-order statistics (spectra) in signal processing and system theory: Theoretical results and some applications. Proceedings of the IEEE, 79(3):278 -- 305, 1991. [55] Fernando Montani, Elena Phoka, Mariela Portesi, and Simon R. Schultz. tical modelling of higher-order correlations in pools of neural activity. 392(14):10.1016/j.physa.2013.03.012, 2013. Statis- Physica A, [56] Ifije E Ohiorhenuan, Ferenc Mechler, Keith P Purpura, Anita M Schmid, Qin Hu, and Jonathan D Victor. Sparse coding and high-order correlations in fine-scale cortical networks. Nature, 466(7306):617 -- 621, July 2010. [57] Jonathan W Pillow, Jonathon Shlens, Liam Paninski, Alexander Sher, Alan M Litke, E J Chichilnisky, and Eero P Simoncelli. Spatio-temporal correlations and visual signalling in a complete neuronal population. Nature, 454(7207):995 -- 999, July 2008. [58] Robert Rosenbaum, James Trousdale, and Kresimir Josi´c. The effects of pooling on spike train correlations. Front Neurosci, 5, 2011. [59] Robert J Rosenbaum, James Trousdale, and Kresimir Josi´c. Pooling and correlated neural activity. Front Comput Neurosci, 4, 2010. [60] Sheldon M. Ross. Stochastic Processes. Wiley, 2 edition, 1 1995. [61] Yasser Roudi, Sheila Nirenberg, and Peter E Latham. Pairwise maximum entropy models for studying large biological systems: when they can work and when they can't. PLoS Comp Biol, 5(5):e1000380, May 2009. [62] E Salinas and T J Sejnowski. Correlated neuronal activity and the flow of neural information. Nat Rev Neurosci, 2(8):539 -- 50, Aug 2001. [63] Elad Schneidman, Michael J Berry, Ronen Segev, and William Bialek. Weak pairwise correla- tions imply strongly correlated network states in a neural population. Nature, 440(7087):1007 -- 1012, April 2006. 25 [64] Elad Schneidman, Micheal J. Berry, Ronen Segev, and William Bialek. Weak pairwise corre- lations imply strongly correlated network states in a neural population. Nature, 440(20):1007 -- 1012, 2006. [65] M. N. Shadlen and W. T. Newsome. The variable discharge of cortical neurons: implications for connectivity, computation, and information coding. J Neurosci, 18:3870 -- 3896, 1998. [66] Maoz Shamir and Haim Sompolinsky. Nonlinear population codes. Neural Comput, 16(6):1105 -- 1136, June 2004. [67] J. Shlens, G.D. Field GD, J.L. Gauthier, M.I. Grivich MI, D. Petrusca D, A. Sher A, A.M. Litke, and E.J. Chichilnisky EJ. The structure of multi-neuron firing patterns in primate retina. J Neurosci, 26:8254 -- 8266, 2006. [68] Jonathon Shlens, Greg D Field, Jeffrey L Gauthier, Martin Greschner, Alexander Sher, Alan M Litke, and EJ Chichilnisky. The structure of large-scale synchronized firing in primate retina. J Neurosci, 29(15):5022 -- 5031, 2009. [69] Wolf Singer. Neuronal synchrony: A versatile code review for the definition of relations? Neuron, 24:49 -- 65, 1999. [70] B. Staude, S. Rotter, and S. Grun. Cubic: cumulant based inference of higher-order correlations in massively parallel spike trains. J Comput Neurosci, 29(1):327 -- 350, 2010. [71] R.L. Stratonovich. Topics in the Theory of Random Noise, volume 2. Gordon and Breach, New York, 1967. [72] A. Tang et al. A maximum entropy model applied to spatial and temporal correlations from cortical networks in vitro. Journal of Neuroscience, 28:505 -- 518, 2008. [73] T. Tetzlaff, S. Rotter, E. Stark, M. Abeles, A. Aertsen, and M. Diesmann. Dependence of neuronal correlations on filter characteristics and marginal spike train statistics. Neural Comp, 20(9):2133 -- 2184, 2008. [74] Tom Tetzlaff, Michael Buschermohle, Theo Geisel, and Markus Diesmann. The spread of rate and correlation in stationary cortical networks. Neurocomputing, 52:949 -- 954, 2003. [75] Simon Thorpe, Arnaud Delorme, and Rufin Van Rullen. Spike-based strategies for rapid processing. Neural Netw, 14(6-7):715 -- 725, 2001. [76] Juan Carlos Vasquez, Olivier Marre, Adrian G Palacios, MJ Berry II, and Bruno Cessac. Gibbs distribution analysis of temporal correlations structure in retina ganglion cells. J Physiology- Paris, 106(3):120 -- 127, 2012. [77] Benjamin White, L F Abbott, and J´ozsef Fiser. Suppression of cortical neural variability is stimulus- and state-dependent. J Neurophysiol, 108(9):2383 -- 2392, November 2012. [78] Ning-long Xu, Mark T Harnett, Stephen R Williams, Daniel Huber, Daniel H O'Connor, Karel Svoboda, and Jeffrey C Magee. Nonlinear dendritic integration of sensory and motor input during an active sensing task. Nature, 492(7428):247 -- 251, December 2012. [79] Shan Yu, Hongdian Yang, Hiroyuki Nakahara, Gustavo S Santos, Danko Nikoli´c, and Dietmar Plenz. Higher-order interactions characterized in cortical activity. J Neurosci, 31(48):17514 -- 17526, 2011. 26 5 Appendix 5.1 Proof of the distributional representation of the GTaS model in Eq. (9) The construction of the GTaS model allows us to provide a useful distributional representation of the process. We describe this representation in a theorem that generalizes Theorem 1 in [11]. This theorem also immediately implies that the GTaS process is marginally Poisson. Some definitions are required: first, for subsets A1, . . . , AN ∈ B(R) and D, D(cid:48) ⊂ D with D ⊂ D(cid:48), let M (D, D(cid:48); A1, . . . , AN ) := B1 × ··· × BN with Bi := Ai, Ac i , R, for i ∈ D, for i ∈ D(cid:48)\D, otherwise In addition, setting 1 = (1, . . . , 1) to be the N -dimensional vector with all components equal to unity, and if QD is a measure on RN , then we define the measure ν(QD) by ν(QD)(A) := for A ∈ B(RN ) QD(A − t1)dt P (Y + t1 ∈ AY ∼ QD)dt. (26) (cid:90) (cid:90) = The measure ν(QD) may be interpreted as giving the expected Lebesgue measure of the subset L of R for which uniform shifts by the elements of L translate a random vector Y ∼ QD in to A. Heuristically, one may imagine sliding the vector Y over the whole real line, and counting the number of times every coordinate ends up in the "right" set -- the projection of A on to that dimension. In equation form, this means ν(QD)(A) = EY[(cid:96)({t ∈ R : Y + t1 ∈ A})Y ∼ QD] . (27) where the subscript Y indicates that we take the average over the distribution of Y ∼ QD. A short proof of this representation is presented below. We now present the theorem, with a proof indicating adjustments necessary to that of [11]. Theorem 0 Let X be an N -dimensional counting process of GTaS type with base rate λ, thinning mechanism p = (pD)D⊂D, and family of shift distributions (QD)D⊂D. Then, for any Borel subsets A1, . . . , AN of the real line, we have the following distributional representation: (cid:88) D(cid:48) 27  X1(A1) ... XN (AN ) (cid:80) (cid:80)   =distr (cid:88)  , D(cid:51)1 ξ(D; A1, . . . , AN ) ... D(cid:51)d ξ(D; A1, . . . , AN ) where the random variables ξ(D; A1, . . . , AN ),∅ (cid:54)= D ⊂ D, are independent and Poisson distributed with E[ξ(D; A1, . . . , AN )] = λ pD(cid:48)ν(QD(cid:48))(M (D, D(cid:48); A1, . . . , AN )). D(cid:48)⊃D Proof. For each marking D(cid:48) ⊂ D, define XD(cid:48) mother process rate λpD(cid:48), shift distribution QD(cid:48), and markings (pD(cid:48) and is zero otherwise (i.e., the only possible marking for XD(cid:48) to be an independent TaS [11] counting process with D = 1 if D = D(cid:48) D )D⊂D where pD(cid:48) is D(cid:48)). We first claim that X =distr XD(cid:48) . (29) (28) To see this, note that spikes in the mother process of the GTaS process of X marked for a set D(cid:48) occur at a rate λpD(cid:48), which is the rate of the process XD(cid:48) . In addition, these event times are then shifted by QD(cid:48), exactly as they are for XD(cid:48) . Thus, the distribution of event times (and hence the counting process distributions) are equivalent. Let A1, . . . , AN be any Borel subsets of the real line. Applying Theorem 1 of [11] to each XD(cid:48) gives the following distributional representation:  , (D; A1, . . . , AN ) ... (D; A1, . . . , AN )  X D(cid:48) ... 1 (A1) X D(cid:48) N (AN )  =distr  (cid:80) D(cid:51)1 ξD(cid:48) (cid:80) D(cid:51)N ξD(cid:48) (cid:105) = λpD(cid:48) (cid:88) D(cid:48)(cid:48)⊃D (cid:104) ξD(cid:48) E (D; A1, . . . , AN ) D(cid:48)(cid:48)ν(QD(cid:48))(M (D, D(cid:48)(cid:48); A1, . . . , AN )) pD(cid:48) = λpD(cid:48)ν(QD(cid:48))(M (D, D(cid:48); A1, . . . , AN )). where the random variables ξD(cid:48) In the latter case, they are independent and Poisson distributed with (D; , A1, . . . , AN ) are taken to be identically zero unless D ⊂ D(cid:48). (30) The second equality above follows from the fact that pD(cid:48) D(cid:48)(cid:48) = 1 if D(cid:48)(cid:48) = D(cid:48) and is zero otherwise. Next, define ξ(D; A1, . . . , AN ) = (cid:88) D(cid:48) (cid:88) D(cid:48)⊃D ξD(cid:48) (D; A1, . . . , AN ) = ξD(cid:48) (D; A1, . . . , AN ). As the sum of independent Poisson variables is again Poisson with rate equal to the sum of the rates, we have that ξ(D; A1, . . . , AN ) is Poisson with mean E[ξ(D; A1, . . . , AN )] = λ pD(cid:48)ν(QD(cid:48))(M (D, D(cid:48); A1, . . . , AN )). (31) Finally, combining Eqs. (29, 30), we may write  X1(A1) ... XN (AN ) D(cid:48)⊃D D(cid:51)1 ξD(cid:48) (cid:88)  =distr  (cid:80) D(cid:48)(cid:80) D(cid:48)(cid:80) (cid:80)  (cid:80) (cid:80) D(cid:51)N ξD(cid:48) D(cid:48) ξD(cid:48) (cid:80) (cid:80)  (cid:80) D(cid:48) ξD(cid:48) (cid:80) D(cid:51)1 = = ... (D; A1, . . . , AN ) ... (D; A1, . . . , AN )  , D(cid:51)N D(cid:51)1 ξ(D; A1, . . . , AN ) D(cid:51)N ξ(D; A1, . . . , AN ) (D; A1, . . . , AN ) ... (D; A1, . . . , AN )  ,  , which, along with Eq. (31), establishes the theorem. A short note: The variable ξ(D; A1, . . . , AN ) counts the number of points which are marked by a set D(cid:48) ⊃ D, but after shifting, only the points attributed to the processes with indices i ∈ D 28 remain in the corresponding subsets Ai. Thus, to determine the number of points attributed to the ith process which lie in Ai (Xi(Ai)), one simply sums the variables ξ for all D containing i, as in Eq. (28). Thus, the intensity of ξ(D; A1, . . . , AN ), λpD(cid:48)ν(QD(cid:48))(M (D, D(cid:48); A1, . . . , AN )), is simply the expected number of such points. Keeping in mind these natural interpretations of terms, Theorem 1 is easier to digest, and the result is not surprising. 5.2 Proof of Eq. (27) In Eq. (27), we gave a more intuitive representation of the measure ν(QD) than the one first defined in [11], which we prove here. Suppose that Q is a measure on B(Rd), and A ∈ B(Rd). Then we have (cid:90) (cid:90)(cid:90) (cid:90)(cid:90) (cid:90) ν(Q)(A) = = = Q(A − t1)dt 1A−t1(y)Q(dy)dt 1{t∈R:y+t1∈A}(t)dtQ(dy) (cid:96)({t ∈ R : y + t1 ∈ A})Q(dy) = = EY[(cid:96)({t ∈ R : Y + t ∈ A})Y ∼ Q] , thus proving Eq. (27) 5.3 Proof of Theorem 1 Theorem 1 Let X be a joint counting process of GTaS type with total intensity λ, marking distribution (pD)D⊂D, and family of shift distributions (QD)D⊂D. Let A1, . . . , Ak be arbitrary sets in B(R), and ¯D = {i1, . . . , ik} ⊂ D with ¯D = k. The cross-cumulant of the counting processes may be written κ(Xi1(A1), . . . , Xik (Ak)) = λ pD(cid:48) P (t1 + Y ¯D ∈ A1 × ··· × AkY ∼ QD(cid:48))dt (32) where Y ¯D represents the projection of the random vector Y on to the dimensions indicated by the members of the set ¯D. Furthermore, assuming that the shift distributions possess densities (qD)D⊂2D, the cross-cumulant density is given by κX i1···ik (τ1, . . . , τk−1) = λ pD(cid:48) ¯D D(cid:48)(t, t + τ1,··· , t + τk−1)dt, q (33) (cid:90) (cid:88) D(cid:48)⊃ ¯D (cid:90) (cid:88) D(cid:48)⊃ ¯D where q ¯D D(cid:48) indicates the kth order joint marginal density of qD(cid:48) in the dimensions of ¯D. Proof. First, as noted in the text, we may rewrite the distributional representation of Theorem 0 (Eq. (28)) as Xi1(Ai1) ...  =distr Xik (Aik )  (cid:80) (cid:80)  i1∈D⊂ ¯D ζD(A1, . . . , AN ) ... ik∈D⊂ ¯D ζD(A1, . . . , AN ) 29 (34) where ζD(A1, . . . , AN ) = (cid:88) D(cid:48)⊃D ( ¯D\D)∩D(cid:48)=∅ ξ(D(cid:48); A1, . . . , AN ). (35) Repeating the description from the main text, the processes ζD are comprised of a sum of all of the processes ξ(D(cid:48)) (defined above, in Theorem 0) such that D(cid:48) contains all of the indices D, but no other indices which are part of the subset ¯D under consideration. These sums are non-overlapping, implying that the ζD are also independent and Poisson. Using the representation of Eq. (34), we first find that κ(Xi1(A1), . . . , Xik (Ak)) = κ =  (cid:88) (cid:88) i1∈D1⊂ ¯D ζD1 , . . . , ··· (cid:88)  (cid:88) ik∈Dk⊂ ¯D ζDk κ[ζD1, . . . , ζDk ]. i1∈D1⊂ ¯D ik∈Dk⊂ ¯D where we suppressed the dependence of the variables ζD on the subsets Ai. The first equality in the previous equation is simply the representation defined in Eq. (35), and the second is from the multilinear property of cumulants (property (C1) in the Methods). Note that the sums are over the sets D1, . . . , Dk satisfying the given conditions. Recall that, by construction, the Poisson processes ζD (see Eq. (35)) are independent for distinct marking sets. Accordingly, the cumulant κ[ζD1, . . . , ζDk ] is zero unless D1 = . . . = Dk, by property (C2) of cumulants -- that is, (cid:40) κk(ζ ¯D(A1, . . . , AN )) Dj = ¯D for each j 0 otherwise . κ[ζD1(A1, . . . , AN ), . . . , ζDk (A1, . . . , AN )] = Hence, κ(Xi1(A1), . . . , Xik (Ak)) = κk(ζ ¯D(A1, . . . , AN )) = E[ζ ¯D(A1, . . . , AN )] , (36) where we have again used that all cumulants of a Poisson-distributed random variable are equal to its mean. For what follows, taking D0, D(cid:48) ⊂ D fixed with D0 ⊂ D(cid:48), the sets M (D, D(cid:48); A1, . . . , AN ) with D0 ⊂ D ⊂ D(cid:48) are disjoint, and (cid:40) Ai, R, i ∈ D0 i /∈ D0 . (37) ∪D0⊂D⊂D(cid:48) M (D, D(cid:48); A1, . . . , AN ) = B1 × ··· × BN with Bi = In particular, note the independence of the above union from D(cid:48). Substituting Eq. (35) in to Eq. (36), we have κ(Xi1(A1), . . . , Xik (Ak)) = D⊃ ¯D (cid:88) (cid:88) (cid:88) (cid:88) (cid:88) D⊃ ¯D D(cid:48)⊃ ¯D D(cid:48)⊃ ¯D D(cid:48)⊃ ¯D = λ = λ = λ E[ξ(D; A1, . . . , Ak] D(cid:48)⊃D pD(cid:48)ν(QD(cid:48))(M (D, D(cid:48); A1, . . . , AN )) (cid:88) pD(cid:48) (cid:88) pD(cid:48)ν(QD(cid:48))(∪ ¯D⊂D⊂D(cid:48)M (D, D(cid:48); A1, . . . , AN )) (cid:90) ν(QD(cid:48))(M (D, D(cid:48); A1, . . . , AN )) ¯D⊂D⊂D(cid:48) = λ pD(cid:48) P (t + Y ¯D ∈ A1 × ··· × AkY ∼ QD(cid:48))dt, 30 where the third equality above is a simple exchange of the order of summation, the fourth equality uses the independence of the inner union from the set D(cid:48) as indicated by Eq. (37), and the final equality follows from the definition of the measure ν(QD(cid:48)) in Eq. (26) and the value of the set union given in Eq. (37). This completes the proof of Eq. (32), and Eq. (33) follows from the definition of the cross- cumulant density in Eq. (24) of the Methods. 5.4 Other details Parameters for figures in the text Figure 1 For figure 1, the GTaS process of size N = 6 consisted of only first order and population- level events which were assigned marking probabilities 0.05 D = D 0.95 6 0 pD = D = {i} for some i ∈ D otherwise . The rate of the mother process was λ = 0.5 kHz, and the shift times for population level events were generated as in Section 2.2 with ϕi ∼ Γ(2, 1) − 1, i = 1, . . . , 6. Figures 3, 4 For figures 3, 4, the GTaS process of size N = 6 consisted of first and second order as well as population-level events. These events had marking probabilities 0.05 D = D 0.95 0 otherwise pD = 21 D = {i},{i, j} for some i, j ∈ D . The rate of the mother process was λ = 0.5 kHz, and the shift times for population level events were generated as in Section 2.2 with ϕi ∼ Exp(0.5), i = 1, . . . , 6. The shift times of the second order events were drawn from an independent Gaussian distribution with each coordinate having standard deviation 5ms. Figure 5 For figure 5, the network parameters were win = 0.4, wsyn = 6, τsyn = 0.1, τd = 1.75. The GTaS input had the same size as the network (N = 10). As in the example of figures 3, 4, the GTaS input included first and second order as well as population level events. Here, we set 0.2 D = D 0.95 0 otherwise pD = 5 D = {i},{i, j} for some i, j ∈ D . The rate of the mother process was λ = 1.5 kHz, and the shift times for population level events were generated as in Section 2.2 with ϕi ∼ Γ(α, β), i = 1, . . . , 6. 31 The shift parameters k, θ (representing shape and scale) were determined by the given shift mean µshift and standard deviation σshift as √ µshift = kθ, σshift = kθ2. The shift times of the second order events were drawn from an independent Gaussian distribution with each coordinate having standard deviation 0.3ms. 5.4.1 Notation table D (pD)D⊂D (QD)D⊂D B(R) D = {1, 2, . . . , N} where N is the system size of the GTaS process under consideration Marking probabilities of a GTaS process. Family of shift distributions on RN for a GTaS process. Borel subsets of the real line R. ξ(D; A1, . . . , AN ) Independent Poisson variables which count points which, after shift- ing, lie in the sets Ai only along the dimensions corresponding to the indices of D. These counts consist of contributions from subsets marked for D(cid:48) ⊃ D, but indices in D(cid:48)\D end up outside the corre- sponding Ai. Defined in the statement of Theorem 0. ζD(A1, . . . , AN ) Independent Poisson variables which are context-dependent resum- mations of the variables ξ(D; A1, . . . , AN ). Defined below Eq. (10). κ(X1, . . . , XN ) Cross-cumulant of the random variables X1, . . . , XN defined in the Methods. κX i1···ik (τ1, . . . , τk−1) Cross-cumulant density defined in Eq. (24). i1···ik−1,pop(τ1, . . . , τk−1) Population cumulant density defined in Eq. (25). κX Table 1: Common notation utilized in the text. 32
1411.7916
4
1411
2016-06-09T14:23:52
The effect of heterogeneity on decorrelation mechanisms in spiking neural networks: a neuromorphic-hardware study
[ "q-bio.NC" ]
High-level brain function such as memory, classification or reasoning can be realized by means of recurrent networks of simplified model neurons. Analog neuromorphic hardware constitutes a fast and energy efficient substrate for the implementation of such neural computing architectures in technical applications and neuroscientific research. The functional performance of neural networks is often critically dependent on the level of correlations in the neural activity. In finite networks, correlations are typically inevitable due to shared presynaptic input. Recent theoretical studies have shown that inhibitory feedback, abundant in biological neural networks, can actively suppress these shared-input correlations and thereby enable neurons to fire nearly independently. For networks of spiking neurons, the decorrelating effect of inhibitory feedback has so far been explicitly demonstrated only for homogeneous networks of neurons with linear sub-threshold dynamics. Theory, however, suggests that the effect is a general phenomenon, present in any system with sufficient inhibitory feedback, irrespective of the details of the network structure or the neuronal and synaptic properties. Here, we investigate the effect of network heterogeneity on correlations in sparse, random networks of inhibitory neurons with non-linear, conductance-based synapses. Emulations of these networks on the analog neuromorphic hardware system Spikey allow us to test the efficiency of decorrelation by inhibitory feedback in the presence of hardware-specific heterogeneities. The configurability of the hardware substrate enables us to modulate the extent of heterogeneity in a systematic manner. We selectively study the effects of shared input and recurrent connections on correlations in membrane potentials and spike trains. Our results confirm ...
q-bio.NC
q-bio
The effect of heterogeneity on decorrelation mechanisms in spiking neural networks: a neuromorphic-hardware study Thomas Pfeil†,1, ∗ Jakob Jordan,2, ∗ Tom Tetzlaff,2 Andreas Grübl,1 Johannes Schemmel,1 Markus Diesmann,2, 3, 4 and Karlheinz Meier1 1Kirchhoff-Institute for Physics, Heidelberg University, Heidelberg, Germany 2Institute of Neuroscience and Medicine (INM-6) and Institute for Advanced Simulation (IAS-6) and JARA BRAIN Institute I, Jülich Research Centre, Jülich, Germany 3Department of Psychiatry, Psychotherapy and Psychosomatics, Medical Faculty, RWTH Aachen University, Aachen, Germany 4Department of Physics, Faculty 1, RWTH Aachen University, Aachen, Germany (Dated: June 10, 2016) 6 1 0 2 n u J 9 ] . C N o i b - q [ 4 v 6 1 9 7 . 1 1 4 1 : v i X r a High-level brain function such as memory, classification or reasoning can be realized by means of recurrent networks of simplified model neurons. Analog neuromorphic hardware constitutes a fast and energy efficient substrate for the implementation of such neural computing architectures in technical applications and neuroscientific research. The functional performance of neural networks is often critically dependent on the level of correlations in the neural activity. In finite networks, correlations are typically inevitable due to shared presynaptic input. Recent theoretical studies have shown that inhibitory feedback, abundant in biological neural networks, can actively suppress these shared-input correlations and thereby enable neurons to fire nearly independently. For networks of spiking neurons, the decorrelating effect of inhibitory feedback has so far been explicitly demon- strated only for homogeneous networks of neurons with linear sub-threshold dynamics. Theory, however, suggests that the effect is a general phenomenon, present in any system with sufficient inhibitory feedback, irrespective of the details of the network structure or the neuronal and synaptic properties. Here, we investigate the effect of network heterogeneity on correlations in sparse, ran- dom networks of inhibitory neurons with non-linear, conductance-based synapses. Emulations of these networks on the analog neuromorphic hardware system Spikey allow us to test the efficiency of decorrelation by inhibitory feedback in the presence of hardware-specific heterogeneities. The configurability of the hardware substrate enables us to modulate the extent of heterogeneity in a systematic manner. We selectively study the effects of shared input and recurrent connections on correlations in membrane potentials and spike trains. Our results confirm that shared-input correla- tions are actively suppressed by inhibitory feedback also in highly heterogeneous networks exhibiting broad, heavy-tailed firing-rate distributions. In line with former studies, cell heterogeneities reduce shared-input correlations. Overall, however, correlations in the recurrent system can increase with the level of heterogeneity as a consequence of diminished effective negative feedback. I. INTRODUCTION Dynamical systems in nature often exhibit a remark- able degree of diversity, specialization or anticorrelation across their components, despite equalizing factors such as common input or homogeneity in component and in- teraction parameters. In many cases, these observations can be explained by the effect of negative feedback. Cell differentiation caused by lateral inhibition [1], formation of new species driven by competition [2] or antiferro- magnetism [3] constitute just a few examples. In recur- rent neuronal networks, inhibitory feedback constitutes a powerful decorrelation mechanism which allows different neurons to respond nearly independently, even if they are driven by largely overlapping local or external inputs [4– 6]. Decorrelation by negative feedback hence implements ∗ These authors contributed equally to this study. † Correspondence: Thomas Pfeil Kirchhoff-Institute for Physics Im Neuenheimer Feld 227 69120 Heidelberg, Germany tel: +49-6221-549813 [email protected] an efficient form of redundancy reduction. In biological systems, it may serve similar purposes as decorrelation in technical applications, where it is used in data compres- sion (e.g., principal-component analysis [7]), crosstalk re- duction (e.g., in digital signal processing [8]), echo sup- pression (e.g., in acoustics [9]) or random-number genera- tion in hardware [10]. Moreover, inhibitory feedback sup- presses "quantization noise" at low frequencies and can thereby increase the dynamical range and signal-to-noise ratio for the encoding of analog signals in the spiking ac- tivity of recurrent neural networks [11]. It is tempting to exploit these mechanisms in synthetic, neurally inspired architectures such as analog neuromorphic hardware. Analog neuromorphic hardware mimics properties of biological neural systems using physical models of neu- rons and synapses (capacitors, for example, emulate in- sulating cell membranes) [12, 13]. The temporal evolu- tion of the analog circuits represents a solution to the corresponding model equations. In consequence, neural- network emulations on analog neuromorphic hardware are massively parallel, extremely fast and energy efficient. Analog neuromorphic devices are therefore highly attrac- tive as tools for neuroscientific research, e.g., for the in- vestigation of learning on long time scales, and technical applications [14–17]. A biologically inspired neural net- work (olfactory system of insects) performing rapid on- line data (odor) classification, for example, has recently been successfully implemented on the analog neuromor- phic hardware system Spikey [18, 19]. In this applica- tion, decorrelation by inhibition is an essential ingredi- ent to guarantee high classification performance. The suppression of quantization noise by inhibitory feedback [11] has been used as a means of noise shaping in several neuromorphic-hardware applications, aiming at the con- struction of biologically inspired ultra-low power analog- to-digital converters [? ? ]. For the functional performance of neuronal architec- tures, the level of correlations between the activities of individual neurons is often pivotal. Whether such cor- relations are beneficial or not is context dependent. A number of previous studies emphasize a functional bene- fit of certain types of correlation for encoding/decoding of information in/from populations of neurons [20–22], in- formation transmission [23–25], robustness against noise [26], or gain control of postsynaptic neurons [27]. Other studies argue that positive cross-correlations are detri- mental as they decrease the precision or sparseness of population codes [19, 28–31]. Cohen & Maunsell [32], for example, have shown that decreased spike-train correla- tions in macaque visual area V4 are accompanied by in- creased behavioral performance in an orientation change- detection task. Depending on the similarity between the trial-averaged responses of different neurons to external stimuli (signal correlation), noise correlations (correla- tions not explained by signal correlations) can either in- crease or decrease the amount of information that can be encoded in or decoded from a population of neurons. In populations of neurons with high signal correlation, van- ishing or even negative noise correlations are desirable to improve the population code [21]. In finite neural networks, an inevitable source of corre- lated neural activity is common presynaptic input, shared by multiple postsynaptic neurons. In network models and in-vivo recordings, however, pairwise correlations in the activity of neighboring neurons have been found to be substantially smaller than expected given the amount of shared input [4, 5, 33–36]. In several studies, this observation has been explained by inhibitory coupling. While Ly et al. [37] and Middleton et al. [38] primarily fo- cused on the effect of feedforward inhibition, Renart et al. [4], Wiechert et al. [39], and Tetzlaff et al. [5] attributed the smallness of correlations to an active decorrelation of neural activity by inhibitory feedback. The mechanism underlying this active decorrelation has already been de- scribed by Mar et al. [11]. In this study, the authors focused on the suppression of low-frequency fluctuations of the population firing rate by recurrent dynamics. As the amplitude of population-rate fluctuations is directly linked to pair-wise correlations (see, e.g., [40]), the effect described in [11] corresponds to a suppression of pairwise correlations in the spiking activity. The theory underly- ing decorrelation by inhibitory feedback suggests the ef- 2 fect to be general: Decorrelation should be observable in any system with sufficiently strong inhibitory feedback, irrespective of the details of the network structure and the cell and synapse properties. For networks of spik- ing neurons, however, the effect has so far been explic- itly demonstrated only for the homogeneous case, where all neurons have identical properties, receive (approxi- mately) the same number of inputs, and, hence, fire at about the same rate [4, 5]. Moreover, the sub-threshold dynamics of individual neurons was assumed to be linear. Biological neuronal networks typically exhibit broad, heavy-tailed firing-rate distributions [41–47], indicating a high degree of heterogeneity, e.g., in synaptic weights [48– 52], in-degrees [53] or time constants [46, 54]. The same holds for neural networks implemented on analog neuro- morphic hardware. All analog circuits suffer from device variations caused by unavoidable variability in the man- ufacturing process. Neurons and synapses implemented in analog neuromorphic hardware therefore exhibit het- erogeneous response properties, similar to their biological counterparts [55, 56]. To understand the dynamics and function of recurrent neural networks in both biological and synthetic substrates, it is therefore essential to ac- count for such heterogeneities. Previous work on recurrent neural networks has shown that heterogeneity in single-neuron properties or connec- tivity broadens the distribution of firing rates [46, 57] and affects the stability of asynchronous or oscillatory states [53, 58–62]. A number of studies pointed at a potential benefit of heterogeneity for the information- processing capabilities of neural networks [62–73]. The effect of heterogeneity on correlations in the activity of recurrent networks of spiking neurons, however, remains unclear. Padmanabhan & Urban [67] have shown that the responses of a population of unconnected neurons are decorrelated by heterogeneity in the neuronal response properties. These results are supported by the subse- quent theoretical analysis in [70]. In the following, we re- fer to this type of decorrelation by heterogeneity as feed- forward decorrelation. It does not account for the effect of the recurrent network dynamics. Active decorrelation due to inhibitory feedback [see above; 4, 5], in contrast, constitutes a very different mechanism. The effect of het- erogeneity on this feedback decorrelation has lately been studied by Bernacchia & Wang [74] in the framework of a recurrent network of linear firing-rate neurons. In this setup, correlations are suppressed by heterogeneity in the network connectivity (distributions of coupling strengths or random dilution of connectivity). It remains unclear, however, whether this holds true for networks of (nonlin- ear) spiking neurons. In this study, we investigate the impact of hetero- geneity on input and output correlations in the asyn- chronous regime of sparse networks of leaky integrate- and-fire (LIF) neurons with conductance-based synapses. Emulation of the networks on the analog neuromorphic hardware system Spikey (Figure 1) [18, 75] enable us to investigate the impact of substrate specific properties on 3 difficult to achieve with hardware studies alone. In anal- ogy to the necessity of performing experiments on biolog- ical neural systems to verify assumptions made in Com- putational Neuroscience, actual emulations on neuromor- phic hardware are essential to understand its properties and develop efficient neural algorithms for these devices. The fact that our main findings hold true for both emu- lations on hardware and simulations with software, and that they can be distilled to simple linear models, sup- port their broad relevance and robustness. II. METHODS A. Network model Details on the network, neuron and synapse model are provided in Table I. Parameter values are given in Ta- ble II. Briefly: We consider a purely inhibitory, sparse network of N (N = 192, unless stated otherwise) LIF neurons with conductance-based synapses. Each neuron receives input from a fixed number K = 15 of randomly chosen presynaptic sources, independently of the network size N. Self-connections and multiple connections be- tween neurons are excluded. Resting potentials El are set above the firing thresholds Θ (equivalent to applying a constant supra-threshold input current). We thereby ensure autonomous firing in the absence of any further external input. Due to temporal noise, the initial condi- tions are essentially random. B. Network emulations on the neuromorphic-hardware system Spikey The Spikey chip (Figure 1) consists of physical models of LIF neurons and conductance-based synapses with ex- ponentially decaying dynamics (for details, see Table I). The emergent dynamics of these physical models repre- sents a solution for the model equations of neurons and synapses in continuous time, in parallel for all units. In contrast, in classical simulations on von-Neumann archi- tectures, model equations are solved by step-wise numer- ical integration, where parallelization is limited by the available number of processor cores. To emphasize the difference between simulations using software and simu- lations using physical models, the term emulation is used for the latter [18]. The response properties of physical neurons and synapses vary across the chip due to unavoidable vari- ations in the production process that manifest in a spa- tially disordered pattern (fixed-pattern noise). In con- trast to the approximately static fixed-pattern noise, temporal noise, including electronic noise and transient experiment conditions (e.g., chip temperature), impairs the reproducibility of emulations. In general, two net- work emulations with identical configuration and stimu- lation do not result in identical network activity. Both The neuromorphic hardware system Spikey. (a) FIG. 1. Photograph of the Spikey chip (size 5 × 5 mm2). It comprises analog circuits of 384 neurons and 98304 synapses, is highly configurable and emulates neural-network dynamics with a speed-up of 104 with respect to biological real-time. (b) Pho- tograph of the partly cased Spikey system, carrying the Spikey chip (covered by a black round seal) and conventional mem- ory. The system is connected to the host computer via USB 2.0, consumes 6 W of power in total and less than 1 nJ per synaptic transmission (see Supplements 1). the network dynamics. Insights about the interplay be- tween features of the computing substrate and network dynamics are a necessary prerequisite for the develop- ment of algorithms that exploit the benefits of analog neuromorphic systems at best. The configurability of this system [18] enables us to systematically vary the level of heterogeneity, and to disentangle the effects of hetero- geneity on feedforward and feedback decorrelation (see above). For simplicity, we focus on purely inhibitory net- works, thereby emphasizing that active decorrelation by inhibitory feedback does not rely on a dynamical balance between excitation and inhibition [5, 6]. We show that decorrelation by inhibitory feedback is effective even in highly heterogeneous networks with broad distributions of firing rates (Section III A). Increasing the level of het- erogeneity has two effects: Feedforward decorrelation is enhanced, feedback decorrelation is impaired. Due to the latter, the overall input and output correlations do not necessarily become smaller with increasing heterogeneity. They can even increase (Section III B). Note that results from specific network emulations on hardware do not directly translate to those obtained by simulations on conventional computers, because the dy- namics, parametrization and interplay of analog circuits is very complex and difficult to reproduce with classical simulations. If simplified models for spatial and tem- poral variability are considered in software simulations, however, emulation results can be reproduced qualita- tively, thereby verifying the design of the hardware sys- tem. While our hardware system is designed to physi- cally implement biologically realistic neural algorithms in a fast and energy-efficient way, software simulations are used as a complementary tool to isolate, verify and in- vestigate different hardware features, such as spatial and temporal parameter variations. Due to the limited access and configurability of network parameters, this would be ab fixed-pattern and temporal noise need to be taken into account when developing models for analog neuromor- phic hardware. The key features of the Spikey chip are the high ac- celeration and configurability of the analog network im- plementation. Some network parameters, e.g., synaptic weights and leak conductances, are configurable for each unit, while other parameters are shared for several units (for details see [18]). The hardware system is optimized for spike in- and output and allows to record the mem- brane potential of one (arbitrarily chosen) neuron with a sampling frequency of 96 MHz in hardware time. On the Spikey chip, capacitances are smaller and conductances are much higher than in biological nervous systems. In consequence, networks on the Spikey chip are emulated with a speed-up of approximately 104 with respect to biological real-time. Due to this high acceleration of the neuromorphic chip, the data bandwidth of the connection between the neuromorphic system and the host computer is not sufficient to communicate with the chip in real time. Consequently, input and output spikes (for stimu- lation and from recordings, respectively) are buffered in a local memory next to the chip. The high acceleration of the Spikey chip allows most of the transistors to operate outside of weak inversion, thereby reducing the effect of transistor variations and minimizing fixed-pattern noise. In contrast to such accelerated systems, most other configurable, analog neuromorphic substrates are de- signed for real-time emulations at very low power con- sumption [76–82] and implement fewer, but more com- plex, neurons [83, 84]. Access to the Spikey system is encapsulated by the simulator-independent language PyNN [85, 86], providing a stable and user-friendly interface. PyNN integrates the hardware into the computational neuroscience tool chain and has facilitated the implementation of several network models on the Spikey chip [18, 19, 87–89]. On the Spikey system, a spiking neural network is emu- lated as follows (Figure 2a): First, the network described in PyNN is mapped to the Spikey chip, i.e., neurons and synapses are allocated and parametrized. Second, input spikes, if available, are prepared on the host computer and transferred to the local memory on the hardware system. Third, the emulation is triggered and available input spikes are generated. Output spikes and mem- brane data are recorded to local memory. Last, spike and membrane data are transferred to the host computer and scaled back into the biological domain of the PyNN model description. For consistency with the model description and sim- plified comparison to the existing literature, all hardware times and all hardware voltages are expressed in terms of the quantities they represent in the neurobiological model, throughout this study. C. Experimental setup 4 To differentiate and compare the effects of shared in- puts and feedback connections on correlations, we inves- tigate two different emulation scenarios: First, we emu- late networks with intact feedback (FB, Figure 2b), and second, the contribution of shared input is isolated by randomizing the temporal order of this feedback (RAND, Figure 2d). In the RAND scenario, the inputs of neurons are de- coupled from their outputs. Spatio-temporal correlations in presynaptic spike trains are removed by randomizing the presynaptic spike times. Input correlations between neurons are measured via their free membrane potential, i.e., the membrane poten- tial with disabled spiking mechanism (technically, the threshold is set very high). Because membrane poten- tial traces can be recorded in the hardware only one at a time, traces are obtained consecutively, while repeat- edly replaying the previously recorded activity of the FB network to a population of unconnected neurons of equal size. We keep the connectivity the same, and hence each neuron receives the same number of spikes as in the recur- rent network during the whole emulation, either without (FBreplay, Figure 2c) or with randomization of presynap- tic spike times (RAND, Figure 2d), respectively. To pre- serve the fixed pattern of variability of synaptic weights in hardware, the same hardware synapses are used for each connection in both scenarios. If network dynam- ics were reproduced perfectly, membrane potential traces and spike times would be identical in the FB and FBreplay cases (see also Section II D). Drawing two different network realizations (i.e., the connectivity matrix) results in the allocation of different hardware synapses, and, due to fixed-pattern noise, in different values of synaptic weights. To average over this variability, throughout this study, emulation results are averaged over M = 100 network realizations, if not stated otherwise. D. Reproducibility of hardware emulations Since the initial conditions of the recurrent network on hardware are undefined, consecutive emulations of the FB network result in different network activities. In the RAND and FBreplay case, however, the input of neurons is decoupled from their output. Although unavoidable temporal noise is present, the system's state-space trajec- tory returns to the trajectory of the previously recorded FB case. A certain degree of reproducibility is required for two reasons: First, the investigated effect of decor- relation by inhibitory feedback requires a precise rela- tion between spike input and output. Thus our method of replacing the feedback loop by replay is only valid if temporal noise does not substantially corrupt this rela- tionship. Second, to record the membrane potentials of all neurons, as if recorded at once, neuron dynamics have 5 FIG. 2. Experimental setup. (a) Data flow of the Spikey system. For details see Section II B. (b) Network with on-chip feedback connections (FB). Spikes from all neurons are recorded to the local memory. (c) Spikes of the FB network in (b) replayed from memory via off-chip spike sources ξi to neurons i (FBreplay). Spike times of ξi correspond to those recorded from neuron i in (b). Spikes from all neurons or the free membrane potential of one selected neuron are recorded. (d) Like (c), but spike times from (b) are randomized for each source (RAND). to be reasonably similar in consecutive emulations. We measure the reproducibility of neuron dynamics by comparing consecutive emulations with identical config- uration, i.e., connectivity and stimulation. For this pur- pose the spiking activity of a FB network is first recorded (Figure 2b) and then repeatedly replayed (Figure 2c). Reproducibility is quantified by the correlations (κX in Table III) of free membrane potential traces and output spike trains obtained for individual neurons in L = 25 different trials. Free membrane potentials are reproduced quite well, while spike trains show larger deviations across trials (Figure 3). Small deviations in the membrane poten- tial (Figure 3b) are amplified by the thresholding proce- dure [39, 90, 91] and can lead to large differences between spike trains (Figure 3c). Consequently, measures based on data of several consecutive replays are more precise for membrane potentials than for spike trains. Nevertheless, results have to be interpreted with care in both cases. E. Calibration The heterogeneity of the Spikey hardware is adjusted by calibrating the leak conductance1 for each individ- ual neuron, compensating for fixed-pattern noise of neu- ron parameters. To this end, a population of uncon- nected neurons is driven by a constant supra-threshold current and the time-averaged population activity ¯r is measured. Then, we applied the bisection method [92] to adjust the leak conductance gl of each neuron, such that the neuron's firing rate matches the target rate ¯r. This results in calibration values b for the leak conductance gl = gl,0(1 + b), where gl,0 is the leak conductance before 1 since capacitances and potentials can not be configured individ- ually for each hardware neuron [18] calibration. Because emulations on hardware are not per- fectly reproducible, more precise calibration was achieved by evaluating the median over 25 identically configured trials instead of single trials. Furthermore, the bisection method was modified for noisy systems (for details, see Supplements 2). Intermediate calibration states are obtained by linearly scaling the full calibration: gl = gl,0(1 + (1 − a)b) . (1) The heterogeneity a is chosen in [0, 1] for calibrations between the uncalibrated (a = 1) and calibrated state (a = 0). In the following, the fully calibrated chip (a = 0) is used, if not stated otherwise. This calibration substantially narrows the distribution of firing rates compared to the uncalibrated state (Fig- ure 4). With respect to the stationary firing rate, vari- ability on the neuron level is reduced from 35.1 to 0.9 s−1. Even in the fully calibrated state, leak conductances can still be widely distributed. Due to the chosen cal- ibration procedure, they are likely to be correlated to other parameters that influence the neurons' response to a constant supra-threshold current after calibration. This mutual compensation can lead to similar phenomenology (here: firing rates) despite disparate parameter values, similar to what is observed in biology [93]. In addition to remaining variations in neuron parameters, synaptic parameters are significantly distributed [19, 94]. F. Correlation measures In the following, we introduce definitions used to ana- lyze the recorded data. For clarity, all relevant equations and their parametrization are listed in Table III and IV, respectively. We quantify correlations of membrane potentials vi(t) and spike trains si(t) by the population-averaged low- 6 FIG. 4. Calibration of the Spikey chip. (a) Histogram of firing rates r for a population of unconnected neurons with supra-threshold input currents, before (gray) and after (black) calibration, each neuron averaged over L = 100 trials. The arrow denotes the target rate ¯r. (b) Difference ∆r = rP 75 − rP 25 of 75th and 25th percentile of the histograms in (a), as a function of network heterogeneity a (Equation 1). The mean firing rate over all values of a is (73.4 ± 0.3) s−1. tials, and output correlations for correlations between spike trains. Shared-input correlations are membrane- potential correlations that are exclusively caused by over- ignoring possible correla- lapping presynaptic sources, tions in the presynaptic activity. In homogeneous net- works, the average pairwise shared-input correlation κV = K N (2) is given by the connectivity K/N [5]. In heterogeneous networks, shared-input correlations can be reduced. In the presence of heterogeneous synaptic weights, for ex- ample, the shared-input correlation κV = 1 1 + CV 2 J K N (3) J ), where CVJ de- is decreased by a factor 1/(1 + CV 2 notes the coefficient of variation of the (non-zero) synap- tic weights. Note, however, that heterogeneities which affect only the spike generation but not the integration of synaptic inputs, e.g. distributions of firing thresholds, have no effect on the shared-input correlation. We assess the significance of correlations by compar- ing the results from emulations to correlations in sur- rogate data, in which we removed spatial correlations. For every neuron, we randomly shuffled bins of the mem- brane potential trace, and assigned a new timestamp uni- formly drawn from the emulation interval to every spike, respectively. We thereby remove all spatio-temporal cor- relations between neurons recorded in parallel. By this procedure we create 100 surrogate trials, across which we calculate the average correlations and the standard error. To quantify fluctuations in the population activity ¯s (Figure 5a–c, horizontal histograms) we compute the power spectrum ¯A(f ) of the population activity (Fig- i (t) and sl FIG. 3. Reproducibility of free membrane potentials and spiking activity in the FBreplay case. (a) Low-frequency co- i (t) and herence κV and κS of free membrane potentials vk i(t), respectively, for i(t) and binned spike trains sk vl each neuron i averaged over L = 25 trials k, l with k (cid:54)= l, for M = 50 different network realizations. The diamond marks the average across all neurons i and M network realizations (b) Free single-trial membrane (κV = 0.96, κS = 0.72). potentials vk i (t) (black), and (c) spike density ξi(t) of a single neuron i for L = 25 identical trials. The selected neuron i has membrane potential coherence and spike train coherence closest to the diamond in (a). i (t) (gray) and average over trials 1 (cid:80)L k=1 vk L frequency coherence κV and κS, respectively. At fre- quency zero, the coherence corresponds to the normal- ized integral of the cross-covariance function, i.e., it mea- sures correlations on all time scales. We define the low- frequency coherence κX, with X ∈ {S, V }, to be the aver- age coherence over a frequency interval from 0.1 to 20 Hz. In this interval, the suppression of population-rate fluc- tuations in recurrent networks due to inhibitory feedback is most pronounced, and the coherence is approximately constant. Before calculating the coherence, we convolve the power- and cross-spectra with a rectangular window to average out random fluctuations. This measure, or a variant of it, is commonly used in the neuroscientific literature [4, 5, 70, 91, 95–97]. We use the terms low- frequency coherence and correlation interchangeably. Throughout this study, the term input correlations is used for correlations between free membrane poten- 0.60.81.0Free membrane potential corr. κV0.20.40.60.8Spike-train corr. κS-74.0-70.0-66.0-62.0-58.0vki(t) (mV)single trialtrial average450046004700480049005000Time t (ms)04812162024Trial index k10-410-310-2Rel. frequencybca050100150Rate r(s−1)050100150200Neuron counta=1a=00.00.51.0Heterogeneity a010203040∆r (s−1)ab 7 FIG. 5. Typical spiking and membrane-potential activity of a random inhibitory network of LIF neurons with intact and cut feedback loop emulated on the fully calibrated system. (a–c) Spiking activity (raster plots), population activity ¯s(t) (horizontal histograms; bin size 50 ms) and time-averaged single-neuron firing rates ri (vertical histograms) in the network with intact feedback (a) and for cases where the feedback loop is cut (b and c). (a) Intact recurrent network (FB scenario). (b) Population of mutually unconnected neurons receiving input spike trains identical to those in (a) (FBreplay scenario). (c) As in (b), but after randomization of presynaptic spike times (RAND scenario). (d,e) Population-averaged cross-correlation functions c(τ ) (after offset subtraction) of pairs of spike trains (d) and power spectra ¯A(f ) (e; log-log representation) of the population activity ¯s(t) (cf. horizontal histograms in (a–c)) for the FB (dark gray), FBreplay (black) and RAND scenario (light gray). Inset in (e): Population-averaged power spectra A(f ) of individual single-cell spike trains (same scales as in main panel). Correlation functions and spectra are averaged across M = 100 network realizations. (f) Membrane potential of a neuron in the RAND scenario (with firing rate of 23.20 s−1 close to population average of 23.24 s−1; see black arrow in (c)) with intact (black curve) and removed threshold (gray curve; free membrane potential). The threshold potential is marked by the horizontal dashed line. The time frame corresponds to the gray-shaded region in (c). ure 5e), which we scale with the duration T of the em- ulation. Consequently, the population power spectrum ¯A(f ), scaled by the population size, coincides with the time-averaged population activity ¯r for high frequencies: limf→∞ 1 N ¯A(f ) = ¯r [35]. As a measure of pairwise correlations in the time do- main (Figure 5d), we compute the population-averaged cross-correlation function c(τ ) by Fourier transforming the population-averaged cross-spectrum C(f ) to time do- main. III. RESULTS In this study, we investigate the roles of shared in- put, feedback and heterogeneity on input and output correlations in random, sparse networks of inhibitory LIF neurons with conductance-based synapses (Table I), im- 182430050100150420043004400450046004700Time t (ms)85756555Mem. potential vi (mV)without thresholdwith threshold200030004000500060007000Time t (ms)1824300501001506040200204060Time lag τ (ms)0.020.010.000.010.02Correlation c(s−1)FBFBreplayRAND100101102Frequency f (Hz)10-1100101102Power spectrum ¯A(s−1)A(s−1)182430¯s(s−1)03060ri(s−1)050100150Neuron index iabcdef plemented on the analog neuromorphic hardware chip Spikey (Figure 1). Similarly to [5], we separate the con- tributions of shared input and feedback by studying dif- ferent network scenarios (Figure 2): In the FB case, we emulate the recurrent network with intact feedback loop (Figure 2b) and record its spiking activity (Figure 5a). In the FBreplay case (Figure 2c), the feedback loop is cut and replaced by the activity recorded in the FB net- work. Ideally, the input to each neuron in the FBreplay case should be identical to the input of the correspond- ing neuron in the FB network. As the replay of spikes and the resulting postsynaptic currents and membrane potentials are not perfectly reproducible on the Spikey chip, the neural responses in the FB and in the FBreplay scenario are slightly different (compare Figures 5a and 5b). In the RAND case (Figures 2d and 5c), we use the same setup as in the FBreplay case. However, the spike times in each presynaptic spike train are random- ized. While the average presynaptic firing rates and the shared-input structure are exactly preserved in this sce- nario, the spatio-temporal correlations in the presynaptic spiking activity are destroyed. Using this setup, we first demonstrate in Section III A that active decorrelation by inhibitory feedback [4, 5] is effective in heterogeneous networks with conductance- base synapses over a range of different network sizes. In Section III B, we show that decreasing the level of het- erogeneity by calibration of hardware neurons leads to an enhancement of this active decorrelation. A. Decorrelation by inhibitory feedback The time-averaged population activities in the FB, FBreplay and RAND scenarios are roughly identical (ver- tical histograms in Figures 5a–c; see also high-frequency power in Figure 5e). In the FB and FBreplay scenario, fluctuations in the population-averaged activity are small (horizontal histograms in Figure 5a and b). The removal of spatial and temporal correlations in the presynaptic spike trains in the RAND case leads to a significant in- crease in the fluctuations of the population-averaged re- sponse activity (horizontal histogram in Figure 5c). At low frequencies (≤ 20 Hz), the population-rate power in the FB and in the RAND case differs by about two orders of magnitudes (dark and light gray curves in Figure 5e). This increase in low-frequency fluctuations in the RAND case is mainly caused by an increase in pairwise correla- tions in the spiking activity (Figure 5d; the power spec- tra of individual spike trains [inset in Figure 5e] are only marginally affected by a randomization of presynaptic spike times) [5]. In other words, shared-input correla- tions, i.e., those leading to large spike-train correlations in the RAND scenario, are efficiently suppressed by the feedback loop in the FB case. On the neuromorphic hardware, the replay of net- work activity is not perfectly reproducible (Section II D). While the across-trial variability in membrane potentials 8 is small, postsynaptic spikes are dithered by few millisec- onds (Figure 3). In the FBreplay case, the suppression of shared-input correlations by correlations in presynaptic spike trains is slightly less efficient as compared to the in- tact network (FB). The differences in the population-rate power spectra and in the spike-train correlations between the FBreplay and RAND case, respectively, are neverthe- less substantial (solid black and light gray curves in Fig- ure 5d and e; note the logarithmic scale; for a detailed investigation of spike dither see Supplements 4 and Sup- plements Figure 7). Note that the suppression of correlations and, hence, population-rate fluctuations by inhibitory feedback is re- stricted to low frequencies (here, to frequencies < 50 Hz; see Figure 5e). In the remainder of this study, we will quantify pairwise correlations by the low-frequency co- herence in the range 0.1–20 Hz (see Section II F). At higher frequencies, the population-rate power spectra in the FB, FBreplay and RAND case are similar. In Fig- ure 5e, the peaks at ∼ 50 Hz and higher harmonics re- sult from the single-cell spike-train statistics (they are also visible in the single-cell spectra; see inset): A large fraction of cells, in particular those firing at higher rates, generate regular spike trains with low inter-spike-interval (ISI) variability (cf. Figure 6d–f). These (fast spiking) cells contribute maxima to the spike-train spectra at fre- quencies close to their firing rates (and higher harmon- ics). The structure of the population-rate spectra at higher frequencies (≥ 50 Hz) is reproduced using surro- gate data where the ISI distributions of the individual neurons (and, hence, their firing rates and ISI variabil- ity) are preserved, but serial ISI correlations and cross- correlations between spike trains are destroyed (data not shown). In the RAND case, presynaptic spike-train correlations free-membrane- were removed, and hence input (i.e., potential) correlations are exclusively determined by the number of shared presynaptic sources (Equation 2). If the in-degree K is fixed, input correlations will decrease with network size N (Equation 2, light gray curve and symbols in Figure 7a). In purely inhibitory networks with intact feedback loop (FB scenario), correlations in presy- naptic spike trains are on average significantly smaller than zero (dark gray diamonds in Figure 7b, [5]), and largely cancel the positive contribution from shared-input correlations. Average input correlations are therefore sig- nificantly reduced (black symbols in Figure 7a). As both shared-input and spike-train correlations scale with the inverse of the network size (N−1; light gray curve in Fig- ure 7a and inset in Figure 7b, respectively) [91], this sup- pression of correlations in the FB (and FBreplay) case is observed for all investigated network sizes N. Note that output correlations are negative even though input cor- relations are positive. This effect is predicted by theory and also observed in linear network models as well as LIF-network simulations on conventional computers (see Figure 9, Supplements 4, Supplements 5 and Section IV). 9 FIG. 6. Modulation of network heterogeneity by leak-conductance calibration (see Section II E). Input (top row) and firing statistics (bottom row) in the intact recurrent networks (FB scenarios) for fully calibrated (a and d; a = 0), partially calibrated (b and e; a = 0.625) and uncalibrated neurons (c and f; a = 1). (a–c) Effect of calibration on input statistics. Distributions of relative mean input D = (¯v − Θ)/σ(v) (distance of time averaged free membrane potential ¯v from firing threshold Θ in units of the standard deviation σ(v)) across the population of neurons. Gray areas in (a), (b) and (c) highlight [−3, 3] intervals, containing 74%, 53% and 42% of the total mass of the distribution, respectively. Inset in (a): Distributions of free membrane potentials v for three neurons α, β and γ with D = −3, D = 0 and D = 3 (arrows in (a)), respectively. Dotted lines mark threshold potentials that may vary due to fixed-pattern noise. (d–f) Effect of calibration on spike-train statistics. Joint (scatter plots) and marginal distributions of single-neuron firing rates r (horizontal histograms; log-linear scale) and coefficients of variation CVISI of inter-spike intervals (vertical histograms; log-linear scale). Dashed lines mark mean of firing rate (22.6 s−1, 28.7 s−1, 34.8 s−1) and CVISI distributions (0.35, 0.28, 0.25), respectively. Gray bars (bottom panels) represent fractions of silent neurons. Data obtained from M = 50 different network realizations. Percentage of dead neurons: 8%, 26%, 37%. FIG. 7. Dependence of population-averaged input correlations (a) and spike-train correlations (b) on the network size N, for the intact network (FB, dark gray diamonds), the FBreplay (black circles) and the RAND (light gray circles) case (fixed in-degree K = 15). Symbols and error bars denote mean and one standard deviation, respectively, across M = 100 network realizations (error bars are partly covered by markers). The gray curve in (a) depicts shared-input correlations in a homogeneous network (Equation 2). The inset in (b) shows a magnified view of the spike-train correlations in the FB case (dark gray diamonds) with a power-law fit ∼ N−1 (dark gray curve). The light gray horizontal band represents mean ± three standard deviations of (spurious) correlations in surrogate data where correlations were removed. Note that free membrane potentials cannot be recorded in the FB case (see Section II). Hence, there are no gray diamonds in (a). 04080120Rate r (s−1)0.00.51.01.52.02.5CVISI0408012004080120201001020Rel. mean input D0.000.050.100.15Rel. neuron countαβγ201001020201001020αβ756555vγdefca=1ba=0.625aa=096112128144160176192Network size N0.000.040.080.12Correlation κSSpike trains96112128144160176192Network size N0.000.040.080.120.16Correlation κVFree membrane potentialsRANDFBreplayFB0.0080.0060.004ab B. Effect of heterogeneity on decorrelation In neural networks implemented in analog neuromor- phic hardware, neuron and synapse parameters vary sig- nificantly across the population of cells (fixed-pattern noise; see Section II B). For a population of mutually unconnected neurons with distributed parameters, injec- tion of a constant (supra-threshold) input current leads to a distribution of response firing rates (Figure 4). In this study, we consider the width of this firing-rate distri- bution as a representation of neuron heterogeneity. It is systematically varied by calibration of leak conductances. The extent of heterogeneity is quantified by the calibra- tion parameter a (a = 1 and a = 0 correspond to the un- calibrated and the fully calibrated system, respectively; for details, see Section II E). For an unconnected popu- lation of neurons subject to constant input, the width of the firing-rate distribution increases monotonically with a. As shown in Figure 6, the level of heterogeneity (i.e., the calibration state a) is clearly reflected in the activ- ity of the recurrent network (FB case). Both the width of the distribution of mean free membrane potentials (Figure 6a–c) as well as the width of the firing-rate dis- tribution increases with a (Figure 6d–f; horizontal his- tograms). In the uncalibrated system (a = 1), a substan- tial fraction of neurons is predominantly driven by con- stant supra-threshold input currents and therefore gen- erates highly regular spike trains (CVISI ≈ 0) with high firing rates (r > 60 s−1). Simultaneously, about 37% of the neurons are silent (r = 0 s−1). Neurons with inter- mediate firing rates (0 s−1 < r < 15 s−1), however, show quite irregular activity (CVISI > 0.5). After calibration, the firing-rate distribution is narrowed. For a = 0, the fraction of silent neurons is reduced to about 8%. Max- imum rates are limited to < 66 s−1. Note that our cal- ibration routine compensates only for the distribution of neuron parameters, but not for the heterogeneity in synapse properties (synaptic weights, synaptic time con- stants; see Section IV). Even for the fully calibrated net- work (a = 0), the firing-rate distribution is therefore still broad. In the RAND case, we obtain similar firing-rate and ISI statistics as in the FB case (Supplements 6). For all levels of heterogeneity attainable by our calibra- tion procedure (a ∈ [0, 1]), input and output correlations are significantly suppressed by the recurrent-network dy- namics (cf. black and dark gray vs. light gray symbols in Figure 8). In a homogeneous, random (Erdős-Rényi) network with fixed in-degree K and linear sub-threshold dynamics, the contribution of shared input to the input (free-membrane-potential) correlation is given by the net- work connectivity K/N [Equation 2; reference 5] (thin light gray curves in Figure 7a and Figure 8a). Nonlinear- ities in synaptic and/or spike-generation dynamics [91] as well as heterogeneity in neuron (and synapse) parameters lead to a suppression of this contribution (Equation 3, [67]). Here, we refer to this type of decorrelation as feed- forward decorrelation. In fact, in our setup the input and 10 spike-train correlations in the RAND case decrease with increasing heterogeneity (light gray symbols in Figure 8). Even in the fully calibrated case input correlations are slightly smaller than K/N (gray symbols vs. thin gray curve in Figure 8a) due to remaining heterogeneities. Spike-train correlations decrease slightly faster with in- creasing heterogeneity than input correlations. These ob- servations indicate that both the synaptic integration and spike generation are affected by heterogeneities on hard- ware. To illustrate the different effects of heterogeneity in synaptic integration and spike generation, we performed network simulations on conventional computers where we distributed either firing thresholds (Figure 9) or synaptic weights (Supplements Figure 4). In the RAND case, the decrease of input correlations on the hardware can be at- tributed to an increase in heterogeneity in parameters af- fecting synaptic integration (compare light gray symbols in Figure 8a and Supplements Figure 4a). Heterogene- ity in spike thresholds, in contrast, does not affect input correlations in the RAND scenario, but strongly reduces spike-train correlations (light gray symbols in Figure 9). Overall, feedforward decorrelation, i.e. the suppression of correlations in the RAND case, becomes more effective in networks with heterogeneous cell parameters. Despite this enhancement of feedforward decorrelation, input and output correlations increase with the level of heterogeneity in the presence of an intact feedback loop (FB and FBreplay scenarios; black and dark gray symbols in Figure 8). We attribute this effect to a weakening of the effective feedback loop in the recurrent circuit: In het- erogeneous networks with broad firing-rate distributions, neurons firing with low or high rates, corresponding to mean inputs far below or far above firing threshold (see Figure 6a–c), are less sensitive to input fluctuations than moderately active neurons (see Supplements Figure 2). Hence, they contribute less to the overall feedback. In consequence, feedback decorrelation is impaired by het- erogeneity (see also Section IV). IV. DISCUSSION We have shown that inhibitory feedback effectively suppresses correlations in heterogeneous recurrent neu- ral networks of leaky integrate-and-fire (LIF) neurons with nonlinear subthreshold dynamics, emulated on ana- log neuromorphic hardware (Spikey; [18, 75]). Both in- put and output correlations are substantially smaller in networks with intact feedback loop (FB), as compared to the case where the feedback is replaced by random- ized input while preserving the connectivity structure and presynaptic firing rates (RAND). Our results hence show that active decorrelation of network activity by in- hibitory feedback [4, 5] is a general phenomenon which can be observed in realistic, highly heterogeneous net- works with nonlinear interaction and sufficiently strong negative feedback. Moreover, the study serves as a proof- of-principle that network activity can be efficiently decor- 11 FIG. 8. Dependence of population-averaged input correlations (a) and spike-train correlations (b) on the heterogeneity of the neuromorphic substrate for the intact network (FB, dark gray diamonds), the FBreplay (black circles) and the RAND (light gray circles) case. Symbols and error bars denote mean and one standard deviation, respectively, across M = 100 network realizations (error bars are partly covered by markers). The gray line in (a) depicts shared-input correlations in a homogeneous network (Equation 2). Insets in (a) and (b) depict magnified views of input correlations in the FBreplay case and spike-train correlations in the FB case, respectively. The light gray horizontal band represents mean ± three standard deviations of (spurious) correlations in surrogate data where correlations were removed. Note that free membrane potentials cannot be recorded in the FB case (see Section II). Hence, there are no gray diamonds in (a). related even on heterogeneous hardware, which can be exploited in functional applications, e.g., in the neuro- morphic algorithms developed by Pfeil et al. [18] and Schmuker et al. [19]. Functional neural architectures often rely on stochas- tic dynamics of its constituents or on some form of back- ground noise [see, e.g., 18, 19, 98]. Deterministic recur- rent neural networks with inhibitory feedback could pro- vide decorrelated noise to such functional networks, both in artificial as well as in biological substrates. In neuro- morphic hardware applications, these "noise networks" could thereby replace conventional random-number gen- erators and avoid a costly transmission of background noise from a host computer to the hardware substrate (which may be particularly relevant for mobile appli- cations with low power consumption; see Supplements 1). It needs to be investigated, however, how well func- tional stochastic circuits perform in the presence of such network-generated noise. Partial calibration of hardware neurons allowed us to modulate the level of network heterogeneity and, there- fore, to systematically study its effect on correlations in the network activity. The analysis revealed two counter- acting contributions: As shown in previous studies [e.g., 67], neuron heterogeneity decorrelates (shared) feedfor- ward input (feedforward decorrelation). On the other hand, however, heterogeneity impairs feedback decorre- lation (see next paragraph). In our network model, this weakening of feedback decorrelation is the dominating factor. Overall, we observed a slight increase in correla- tions with increasing level of heterogeneity. We cannot exclude that feedforward decorrelation may play a more significant role for different network configurations (e.g., different connection strengths or network topologies, dif- ferent structure of external inputs, different types of het- erogeneity). Our study demonstrates, however, that het- erogeneity is not necessarily suppressing correlations in recurrent systems. In this context, it would be interest- ing to investigate the interplay of signal and noise corre- lations in the presence of network heterogeneities in re- current systems. We leave this intriguing topic to future studies. As shown in [5], feedback decorrelation in recurrent networks becomes more (less) efficient with increasing (decreasing) strength of the effective negative feedback. For networks of spiking neurons, the effective connection strength wij between two neurons j and i corresponds to the total number of extra spikes emitted by neuron i in re- sponse to an additional input spike generated by neuron j (see, e.g., [99]). Assuming that the effect of a single addi- tional input spike is small, the effective connectivity can be obtained by linear-response theory [91]. Note that the effective weights wij depend on the working point, i.e., the average firing rates of all pre- and postsynaptic neu- rons (mathematically, wij is given by the derivative of the stationary response firing rate ri = φi(r1, . . . , rj, . . . , rN ) of neuron i with respect to the input firing rate rj, eval- uated at the working point; for details, see [5]). Neurons firing at very low or very high rates are typically less sen- sitive to input fluctuations than neurons firing at inter- mediate rates (due to the shape of the response function φi(r1, . . . , rN )). Their dynamical range is reduced. In consequence, they hardly mediate feedback in a recurrent network. In heterogeneous networks with broad distribu- tions of firing rates, the number of these insensitive neu- rons is increased. Hence, the effective feedback is weak- 0.00.20.40.60.81.0Heterogeneity a0.010.000.010.020.030.040.05Correlation κSSpike trains0.00.20.40.60.81.0Heterogeneity a0.000.020.040.060.08Correlation κVFree membrane potentialsRANDFBreplayFB0.00380.00320.0050.0100.015ab ened (see Supplements 3). We can qualitatively repro- duce this effect of heterogeneity on correlations in recur- rent networks (FB case) by means of a simplified linear rate model where increasing heterogeneity is described as a decrease in the effective-weight amplitudes (see Supple- ments 5). A more quantitative analysis requires an ex- plicit mapping of the synaptic weights in the LIF-neuron network to the effective weights of the linear model (as in, e.g., [100]) in the presence of distributed firing rates. We commit this task to future studies. Note that the rate dependence of the effective weights and the resulting ef- fects on correlations are consistent with our observation that neuron pairs with very low firing rates exhibit spike- train correlations close to zero, whereas pairs with high firing rates are positively correlated (see Supplements 7). Pairs with one neuron firing at an intermediate rate of- ten exhibit negative spike-train correlations. As shown in [4, 5], these negative spike-train correlations are essen- tial for compensating the positive contribution of shared inputs to the total input correlation (at least in purely in- hibitory networks). Narrowing the firing rate distribution (e.g., by calibration of hardware neurons) increases the number of neurons contributing to the negative feedback, which, in turn, leads to more neuron pairs with negative spike-train correlations and, therefore, to smaller overall correlations. Seemingly contrary to our findings, Bernacchia & Wang [74] report a decrease in correlations with increas- ing level of heterogeneity. The results of their study are obtained for a linear network model, which can be consid- ered the outcome of the linearization procedure described above. Hence, the connectivity of their model corre- sponds to an effective connectivity (see above). Their study neglects the rate (working-point) dependence of the effective weights and can therefore not account for the ef- fect of firing-rate heterogeneity. In [74], heterogeneity is quantified by the variance of the (effective) weight matrix (Equations 2.2 and 2.4 in [74]). For sparse connectivity matrices (with a large number of zero elements), the vari- ance of the weight matrix reflects not only the width of the non-zero-weight distribution, but also its mean (Equation 2.4 in [74]). For networks of nonlinear spik- ing neurons, heterogeneities in neuron and/or synapse parameters broadens the distribution of non-zero effec- tive weights, but may simultaneously reduce its mean (see above, Supplements 3, and [46, 100]). Hence, the variance of the full weight matrix may decrease (for il- lustration, see Supplements Figure 9). In other words, increasing heterogeneity in the nonlinear system may cor- respond to decreasing heterogeneity in the linearized sys- tem. A direct test of this hypothesis requires an explicit linearization of the nonlinear heterogeneous system. The results of this study were obtained by network emulations on analog neuromorphic hardware. We re- produced the main findings by means of conventional computer simulations of LIF-neuron networks with dis- tributed firing thresholds (see Figure 9). The focus on threshold heterogeneity allows us to isolate the effect 12 of firing-rate distributions on correlations. It does not affect shared-input correlations (see Section II F). Al- though networks simulated on conventional computers and those emulated on the neuromorphic hardware dif- fer in several respects (e.g., in the exact implementation of heterogeneity or the synapse model; compare Supple- ments Table I and II to Table I and II, respectively), the qualitative results are very similar: In networks with intact feedback loop, input and output correlations are substantially reduced (as compared to the case where the feedback is replaced by randomized input), but increase with the extent of heterogeneity. As predicted by the theory for homogeneous inhibitory networks, we observe positive input correlations and negative output correla- tions (see Equation 21 in [5] and in the paragraph which follows; see also [103] and Supplements 5). Further, note that heterogeneity in neuron parameters does not "av- erage out" in larger networks. Upscaling the network size by a factor of 25 (N = 4800, in-degree K = 375) yields smaller spike-train correlations, but the qualita- tive results are similar to those obtained for the smaller network (N = 192, K = 15) emulated on the Spikey chip (compare Figure 8 to Supplements Figure 3). In networks with intact feedback loop (FB and FBreplay scenarios), the precise spatio-temporal structure of spike trains arranges such that the self-consistent input and output correlations are suppressed. Perturbations of this structure in the local input typically lead to an increase in correlations [5]. In this study, we demonstrate this by replaying spiking activity after randomization of spike times, i.e., by replacing the time of each input spike by a random number uniformly drawn from the full emulation time interval [0, T ) (RAND case). However, even subtle modifications of input spike trains, such as random dither of spike times by few milliseconds, lead to an increase of correlations. On the neuromorphic hardware, replay of spike trains is not entirely reproducible (see Section II D and Supplements 9). Hence, spike-train correlations mea- sured in the FBreplay mode are slightly larger than in the FB case. We would expect the same effect on the in- put side (free membrane potentials). Due to hardware limitations, however, we can measure input correlations only in replay mode (FBreplay or RAND), but not in the fully connected network (FB). Therefore, all reported in- put correlations are likely to be slightly overestimated. In conventional network simulations, we mimicked the effect of unreliable replay by input-spike dithering and, indeed, find a gradual increase in input and output cor- relations (see Supplements Figure 6). These results seem to be contrary to the study by Rosenbaum & Josic [104], in which synaptic noise leads to a decrease of output correlations in a feedforward scenario. In our case, spike- train correlations, which suppress shared-input correla- tions, are removed by dithering spikes, thereby increas- ing correlations on the output side. In [104], in contrast, spike-train correlations are always zero, and shared-input correlations are decreased by synaptic failure, explaining the decreased output correlations. We attribute this con- 13 FIG. 9. Dependence of population-averaged input correlations (a), and spike-train correlations (b) on the width of threshold distributions in networks simulated with NEST [101] and PyNN [102], for the intact network (FB, dark gray circles) and the RAND (light gray circles) case. Symbols and error bars denote mean and standard deviation, respectively, across M = 30 network realizations (error bars are partly covered by markers). The gray line in (a) depicts shared-input correlations in a homogeneous network (Equation 2). The inset in (b) shows a magnified view of the spike-train correlations in the FB case. Note that in simulations the FBreplay is identical to the FB case, and is hence not shown. For details see Supplements 4. tradiction to the missing feedback loop in their system, and expect correlations to increase in recurrent networks subject to similar perturbations. Despite the imperfect replay of input spikes, the decor- relation effect is clearly visible in hardware emulations, both on the input and on the output side. The repro- ducibility of emulations on neuromorphic hardware could be improved by stabilizing the environment of the sys- tem, e.g., the chip temperature or the support electron- ics (under development). Analog hardware, however, will never reach the level of reproducibility of digital com- puters. But note that, similar to analog hardware, bi- ological neurons exhibit a considerable amount of trial- to-trial variability, even under controlled in-vitro condi- tions [90]. So far, the details of how neuronal noise, for example, stochastic synapses (spontaneous postsynap- tic events, stochastic spike transmission, synaptic failure [105]), affects correlations in recurrent neural circuits re- main unclear. Although different Spikey chips exhibit different real- izations of fixed-pattern noise, they show a comparable extent of heterogeneity and yield results which are qual- itatively similar to those presented in this article (Sup- plements 8). We have shown that negative feedback in recurrent cir- cuits can efficiently suppress correlations, even in highly heterogeneous systems such as the analog neuromorphic architecture Spikey. Correlations can be further reduced by minimizing the level of network heterogeneity. In this study, we reduced the level of heterogeneity through cal- ibration of neuron parameters in the unconnected case (see Section II E). The calibration could, in principle, be improved by calibrating neuron (and possibly synapse) parameters in the full recurrent network. Such calibra- tion procedures are however time consuming and cumber- some. In biological substrates, homeostasis mechanisms [56, 106] keep neurons in a responsive regime and reduce the level of firing-rate heterogeneity in a self-regulating manner. Future neuromorphic devices could mimic this behavior, thereby reducing the necessity of time consum- ing calibration procedures. Alternatively, the analog cir- cuits could be optimized to reduce fixed-pattern noise. This would likely require the allocation of more chip re- sources, hence reducing the network size per chip area. For simplicity, this work focuses on purely inhibitory networks (as in [11]). This demonstrates that decorrela- tion by inhibitory feedback does not rely on a dynami- cal balance between excitation and inhibition (note that the external "excitatory" drive is constant in our model) [5, 6]. Previous studies have shown that, for the ho- mogeneous case, decorrelation by inhibitory feedback is a general phenomenon, which also occurs in excitatory- inhibitory networks, provided the overall inhibition is sufficiently strong (which is typically the case to ensure stability) [4–6, 74]. For the heterogeneous case, com- puter simulations of excitatory-inhibitory networks show qualitatively the same results as purely inhibitory net- works (compare Figure 8 to Supplements Figure 5), con- firming that our results generalize to the case of mixed excitatory-inhibitory coupling. Similar to our study, Giulioni et al. [107] use a theory- guided approach to implement, verify and investigate network dynamics on analog neuromorphic hardware. In their study, an attractor network is implemented that is inspired by a mean-field model. Due to heterogeneities in synaptic efficacies on the hardware, stability analy- sis of attractor states requires the authors to measure effective response functions of populations of hardware neurons. To this end, they replace recurrent connections in one population of neurons by external input. This al- 0.000.020.040.060.080.100.120.14Rel. std. deviation of threshold potential σΘµΘ0.010.000.010.020.030.040.05Correlation κS0.000.020.040.060.080.100.120.14Rel. std. deviation of threshold potential σΘµΘ0.000.020.040.060.080.10Correlation κVFBRAND0.0060.0030.000ab 14 mized by processing the data directly on the hardware, or by choosing a data representation which is closer to the format used on the Spikey chip, but this would impair user-friendliness, and hence, the effectiveness of proto- typing. While the Spikey system permits the monitoring of the spiking activity of all neurons simultaneously, ac- cess to the membrane potentials is limited to a single (al- beit arbitrary) neuron in each emulation run. Monitor- ing of membrane potentials of a population of n neurons therefore requires n repetitions of the same emulation. Extending the hardware system to enable access to the membrane potentials of at least two neurons simultane- ously would allow for a direct observation of input cor- relations in the intact network (and thereby avoid prob- lems with replay reproducibility; see above) and reduce execution time (the Spikey chip itself permits recording of up to eight neurons in parallel, the support electron- ics, however, does not). While the Spikey system does not significantly outperform conventional computers in terms of computational power, emulations on this sys- tem are much more energy efficient (Supplements 1). A substantial increase of computational power is expected for large systems exploiting the scalability of this tech- nology without slow-down [108]. ACKNOWLEDGMENTS We would like to thank Matthias Hock, Andreas Har- tel, Eric Müller and Christoph Koke for their support in developing and building the Spikey system, as well as Mihai Petrovici and Sebastian Schmitt for fruitful dis- cussions. This research is partially supported by the Helmholtz Association portfolio theme SMHB, the Jülich Aachen Research Alliance (JARA), EU Grant 269921 (BrainScaleS), and EU Grant 604102 (Human Brain Project, HBP). We acknowledge financial support by Deutsche Forschungsgemeinschaft and Ruprecht-Karls- Universität Heidelberg within the funding programme Open Access Publishing. All network simulations car- ried out with NEST (http://www.nest-simulator.org). Appendix A: Network description Details on the network model as well as parameter val- ues are provided in Table I and II, respectively. Appendix B: Description of data analysis A summary of the quantities and the data analysis used in this study is provided in Table III. Parameter values for the data analysis are given in Table IV. FIG. 10. Acceleration factor as a function of emulated net- work time T for the record (black) and the replay case (gray). The acceleration factor is defined as the ratio between the emulated network time T (in biological time) and the exe- cution time (wall clock time). In the record case, a network realization is generated on the host computer and uploaded to the chip. During the subsequent emulation, spike trains are recorded. In the replay case, spikes are replayed and the membrane potential of one neuron is recorded with full sam- pling frequency (9.6 kHz). The execution time covers the full data flow from a network description in PyNN to the emula- tion on the Spikey system and back to the network repre- sentation in PyNN. The time-averaged population firing rate is ¯r = (23.7 ± 0.2) s−1. The vertical dashed line depicts the runtime used in this study. The hardware system has to be initialized once before usage (< 1 s), which is not considered here. lows them to measure the firing rate of the population as a function of the external input, while the activity of the population is in equilibrium with that of other re- currently connected populations. This study represents another example illustrating that investigations of actual hardware emulations are a prerequisite for successful ap- plication of analog neuromorphic hardware. This study demonstrates that the Spikey system has matured to a level that permits its use as a tool for neu- roscientific research. For the results presented in this study, we recorded in total 1011 membrane-potential and spike-train samples, representing more than 100 days of biological time. Due to the 104-fold acceleration of the Spikey chip, this corresponds to less than 15 minutes in the hardware-time domain. Interfacing the hardware system, however, reduces the acceleration to an approx- imately 50-fold speed-up (Figure 10). The translation between the network description and its hardware rep- resentation claims the majority of execution time, more than the network emulation and the transfer of data to and from the hardware system together. Encoding and decoding spike times on the host computer is particu- larly expensive. Obviously, the system could be opti- 100101102103Emulated network time T(s)100101102103Acceleration factorrecordreplay 15 A Populations Topology Connectivity Neuron model Channel models Synapse model Plasticity External input Measurements Other B Name I Model summary One (inhibitory) None Random convergent connections (fixed in-degree) Leaky integrate-and-fire (LIF), fixed firing threshold, fixed absolute refractory time None Exponentially decaying conductances, fixed delays None Resting potential higher than threshold (= constant current) (El > Θ) Spikes and membrane potentials from all neurons No autapses, no multapses Populations Elements LIF neuron Size N C Source Target Pattern I I Random convergent connections, in-degree K Connectivity D Type Subthreshold dynamics Conductance dynamics Spiking Neuron and synapse model d Cm dt v(t) = −gl(v(t) − El) − gsyn(t)(v(t) − Einh) Leaky integrate-and-fire, exponentially decaying conductances Subthreshold dynamics (t (cid:54)∈ (t∗, t∗ + τref)): Reset and refractoriness (t ∈ (t∗, t∗ + τref)): This model is emulated by analog circuitry on the Spikey chip [13]. For each presynaptic spike at time t∗ (t > t∗ + d): v(t) = vreset gsyn(t) ≈ J exp(− t−t∗−d τsyn )Θ(t), with J = whwgmax and Heaviside function Θ(t). This model is emulated by analog circuitry on the Spikey chip [18]. If v(t∗−) < Θ ∧ v(t∗+) ≥ Θ: emit spike with time stamp t∗ TABLE I. Description of the network model (according to [109]). [1] Gierer, A., & Meinhardt, H. (1974). Biological pat- tern formation involving lateral inhibition. Lectures on mathematics in the life sciences 7, 163. [2] Rosenzweig, M. L. (1978). Competitive speciation. Bi- ological Journal of the Linnean Society 10 (3), 275–289. [3] Neel, L. (1952). Antiferromagnetism and ferrimag- netism. Proceedings of the Physical Society. Section A 65 (11), 869. [4] Renart, A., De La Rocha, J., Bartho, P., Hollender, L., Parga, N., Reyes, A., & Harris, K. D. (2010). The asyn- chronous state in cortical circuits. Science 327, 587–590. [5] Tetzlaff, T., Helias, M., Einevoll, G., & Diesmann, M. (2012). Decorrelation of neural-network activity by in- hibitory feedback. PLoS Comput. Biol. 8 (8), e1002596. [6] Helias, M., Tetzlaff, T., & Diesmann, M. (2014). The correlation structure of local cortical networks intrinsi- cally results from recurrent dynamics. PLoS Comput. Biol. 10 (1), e1003428. [7] Pearson, K. (1901). Liii. on lines and planes of closest fit to systems of points in space. Philosophical Magazine Series 6 2 (11), 559–572. B Name N C Name K D Name vreset El Θ Einh τref Cm gl τ eff m D Name gmax whw V eff max τ eff syn deff Other Name SpikeyHAL PyNN vmodule logger Other Name 16 Values {96, 112, 128, 144, 160, 176, 192} Populations Description network size Connectivity Values 15 Values −80 mV −52 mV −62 mV −80 mV 1 ms 0.2 nF tuned during calibration process uncalibrated: 10.50(5.94, 17.90)ms calibrated: 10.77(6.22, 18.30)ms Description number of presynaptic partners Neuron Description reset potential resting potential firing threshold inhibitory reversal potential refractory period membrane capacitance leak conductance effective membrane time constant Synapse Description conductance amplitude synaptic weight (in hardware values ∈ [0, 15]) Values in the order of 1 nS 3 uncalibrated: -5.57(−12.50,−2.62)mV effective post-synaptic potential prefactor calibrated: -6.05(−14.37,−2.92)mV uncalibrated: 3.78(2.52, 5.58)ms calibrated: 4.17(2.76, 6.18)ms uncalibrated: 2.16(1.91, 2.48)ms calibrated: 2.26(1.93, 2.55) effective synaptic-current time constant effective synaptic delay Values 9e86d11c 2fe40b43 76ef3b44 826c5ed6 Software Description git revision git revision git revision git revision Hardware Values chip 508 (version 5) chip 503, 504 (version 5) and 603, 605, 666 (version 4) Description chip used for main manuscript chips used for supplements TABLE II. Parameter values for the network model described in Table I. Bold numbers indicate default values. Gray numbers indicate target values not considering fixed-pattern noise. Leak conductances gl are adjusted in the calibration process (see Section II E). τ eff max describe effective values measured from spike-triggered averages in RAND emulations as described in Supplements 10. Effective values denote the median across synapses and trials followed by the 25th and 75th percentiles in brackets. syn, deff and V eff m , τ eff A Measure spike density spike train population activity time-averaged population activity membrane potential (finite time) Fourier transform (single unit) power spectrum population-averaged power spectrum population power spectrum pairwise cross spectrum population-averaged cross spectrum sliding window filter coherence low-frequency coherence pop.-averaged cross-correlation function time average 17 Analysis measures Details ξi(t) =(cid:80) k δ(t − tk i ) (cid:80) si(tk) = number of spikes of neuron i per bin [k∆t, (k + 1)∆t) ¯s(t) = 1 N ¯r = (cid:104)¯s(t)(cid:105)t vi(tk) = membrane potential of neuron i in bin [k∆t, (k + 1)∆t) i si(t) 0 dt xi(t)e−2πif t (with inverse F−1) j Sj(f )) (cid:80) T X∗ T ((cid:80) Xi(f ) = F[xi(t)](f ) =(cid:82) T i (f )Xi(f ) i (f ))((cid:80) i Ai(f ) i S∗ (cid:80) i(cid:54)=j Cij(f ) ≡ Ai(f ) = 1 A(f ) = 1 N ¯A(f ) = 1 i (f )Xj(f ), i (cid:54)= j T X∗ Cij = 1 C(f ) = N (N−1) (note: C(f ) ∈ R) X(f ) → X(f ) ∗ H(f ) with H(f ) = 1 κ(f ) = C(f ) A(f ) 1 fmax−fmin N (N−1) f1−f0 κX = 1 1 c(τ ) = (cid:104). . .(cid:105)t N (N−1) ( ¯A(f ) − N A(f )) 1 Θ(f − f0)Θ(f1 − f ) (cid:82) fmax (cid:80) i(cid:54)=j (cid:104)si(t)sj(t + τ )(cid:105)t ≡ F−1[C(f )](τ ) df κ(f ) fmin TABLE III. Summary of the data analysis. Here i ∈ [1, N ], X ∈ S, V . A Parameter ∆t ∆tm Twarmup T M ∆F fmin, fmax a Analysis parameters Description bin size for spike trains bin size for membrane potential traces initial warm-up time (not considered in analysis) emulated network time (biol. time domain) number of network realizations width of sliding window interval boundaries for low-frequency coherence calibration state Values 1 ms 0.52 ms 1 s 10 s {50, 100} 1 Hz 0.1 Hz, 20 Hz {0, 1 8 , 4 8 , 5 8 , 3 8 , 2 8 , 6 8 , 7 8 , 1} TABLE IV. Summary of analysis parameters (default values in bold). [8] Ginis, G., & Peng, C.-N. (2006). Alien crosstalk cancellation for multipair digital subscriber line sys- tems. EURASIP Journal on Advances in Signal Pro- cessing 2006 (1), 1–12. [9] Benesty, J., Gänsler, T., Morgan, D. R., Sondhi, M. M., & Gay, S. L. (2001). Advances in Network and Acoustic Echo Cancellation (1 ed.). Berlin, Germany: Springer. [10] Trichina, E., Bucci, M., De Seta, D., & Luzzi, R. (2001). Supplemental cryptographic hardware for smart cards. Micro, IEEE 21 (6), 26–35. [11] Mar, D., Chow, C., Gerstner, W., Adams, R., & Collins, J. (1999). Noise shaping in populations of coupled model neurons. Proc. Nat. Acad. Sci. USA 96 (18), 10450– 10455. [12] Mead, C. (1990). Neuromorphic electronic systems. Proc. IEEE 78 (10), 1629–1636. [13] Indiveri, G., Linares-Barranco, B., Hamilton, T. J., van Schaik, A., Etienne-Cummings, R., Delbruck, T., Liu, S.-C., Dudek, P., Häfliger, P., Renaud, S., Schemmel, J., Cauwenberghs, G., Arthur, J., Hynna, K., Folowosele, F., Saighi, S., Serrano-Gotarredona, T., Wijekoon, J., Wang, Y., & Boahen, K. (2011). Neuromorphic silicon neuron circuits. Front. Neurosci. 5 (73). [14] SenseMaker (2009). SenseMaker: A multi-sensory, task- specific, adaptable perception system; project website. Available at: http://cordis.europa.eu/project/rcn/ 62746_en.html. [15] FACETS (2010). Fast Analog Computing with Emer- gent Transient States, project website. Available at: http://www.facets-project.org. [16] BrainScaleS (2014). Project website. Available at: http://www.brainscales.eu. [17] Human Brain Project (2014). Project website. Available at: http://www.humanbrainproject.eu. [18] Pfeil, T., Grübl, A., Jeltsch, S., Müller, E., Müller, P., Petrovici, M. A., Schmuker, M., Brüderle, D., Schem- mel, J., & Meier, K. (2013). Six networks on a uni- versal neuromorphic computing substrate. Front. Neu- rosci. 7 (11). [19] Schmuker, M., Pfeil, T., & Nawrot, M. P. (2014). A neu- romorphic network for generic multivariate data classi- fication. Proc. Natl. Acad. Sci. USA 111 (6), 2081–2086. [20] Ecker, A. S., Berens, P., Tolias, A. S., & Bethge, M. (2011). The effect of noise correlations in populations of diversely tuned neurons. J. Neurosci. 31 (40), 14272– 14283. [21] Averbeck, B. B., Latham, P. E., & Pouget, A. (2006). Neural correlations, population coding and computa- tion. Nat. Rev. Neurosci. 7, 358–366. [22] Moreno-Bote, R., Beck, J., Kanitscheider, I., Pitkow, Information- X., Latham, P., & Pouget, A. (2014). limiting correlations. Nat. Neurosci. 17, 1410–1417. [23] Fries, P. (2005). A mechanism for cognitive dynamics: neuronal communication through neuronal coherence. Theor. Comput. Sci. 9 (10), 474–480. [24] Abeles, M. (1991). Corticonics: Neural Circuits of the Cerebral Cortex (1st ed.). Cambridge: Cambridge Uni- versity Press. [25] Diesmann, M., Gewaltig, M.-O., & Aertsen, A. (1999). Stable propagation of synchronous spiking in cortical neural networks. Nature 402 (6761), 529–533. [26] Tabareau, N., Slotine, J.-J., & Pham, Q.-C. (2010). How synchronization protects from noise. PLoS Com- put. Biol. 6 (1). doi:10.1371/journal.pcbi.1000637. 18 [27] Salinas, E., & Sejnowski, T. J. (2001). Correlated neu- ronal activity and the flow of neural information. Nat. Rev. Neurosci. 2 (8), 539–550. [28] Shadlen, M. N., & Newsome, W. T. (1998). The variable discharge of cortical neurons: Implications for connec- tivity, computation, and information coding. J. Neu- rosci. 18 (10), 3870–3896. [29] Abbott, L. F., & Dayan, P. (1999). The effect of corre- lated variability on the accuracy of a population code. Neural Comput. 11, 91–101. [30] Tripp, B., & Eliasmith, C. (2007). Neural popula- tions can induce reliable postsynaptic currents without observable spike rate changes or precise spike timing. Cereb. Cortex 17 (8), 1830–1840. [31] Schmuker, M., & Schneider, G. (2007). Processing and classification of chemical data inspired by insect olfac- tion. Proc. Natl. Acad. Sci. USA 104 (51), 20285–20289. [32] Cohen, M. R., & Maunsell, J. H. R. (2009). Attention improves performance primarily by reducing interneu- ronal correlations. Nat. Neurosci. 12, 1594–1600. [33] Tetzlaff, T., Morrison, A., Geisel, T., & Diesmann, M. (2004). Consequences of realistic network size on the stability of embedded synfire chains. Neurocomput- ing 58–60, 117–121. [34] Kriener, B., Tetzlaff, T., Aertsen, A., Diesmann, M., & Rotter, S. (2008). Correlations and population dynam- ics in cortical networks. Neural Comput. 20, 2185–2226. [35] Tetzlaff, T., Rotter, S., Stark, E., Abeles, M., Aertsen, A., & Diesmann, M. (2008). Dependence of neuronal correlations on filter characteristics and marginal spike- train statistics. Neural Comput. 20 (9), 2133–2184. [36] Ecker, A. S., Berens, P., Keliris, G. A., Bethge, M., & Logothetis, N. K. (2010). Decorrelated neuronal firing in cortical microcircuits. Science 327 (5965), 584–587. [37] Ly, C., Middleton, J., & Doiron, B. (2012). Cellular and circuit mechanisms maintain low spike co-variability and enhance population coding in somatosensory cor- tex. Front. Comput. Neurosci. 6 (7), 1–25. [38] Middleton, J., Omar, C., Doiron, B., & Simons, D. (2012). Neural correlation is stimulus modulated by feedforward inhibitory circuitry. J. Neurosci. 32 (2), 506–518. [39] Wiechert, M. T., Judkewitz, B., Riecke, H., & Friedrich, R. W. (2010). Mechanisms of pattern decorrelation by recurrent neuronal circuits. Nat. Neurosci. 13, 1003– 1010. [40] Harris, K. D., & Thiele, A. (2011). Cortical state and attention. Nat. Rev. Neurosci. 12, 509–523. [41] Griffith, J. S., & Horn, G. (1966). An analysis of spon- taneous impulse activity of units in the striate cortex of unrestrained cats. J. Physiol. (Lond.) 186 (3), 516–534. [42] Koch, K. W., & Fuster, J. M. (1989). Unit activity in monkey parietal cortex related to haptic perception and temporary memory. Exp. Brain Res. 76 (2), 292–306. [43] Shafi, M., Zhou, Y., Quintana, J., Chow, C., Fuster, J., & Bodner, M. (2007). Variability in neuronal ac- tivity in primate cortex during working memory tasks. Neuroscience 146 (3), 1082–1108. [44] Hromadka, T., DeWeese, M. R., & Zador, A. M. (2008). Sparse representation of sounds in the unanesthetized auditory cortex. PLoS Biol. 6 (1). [45] OĆonnor, D. H., Peron, S. P., Huber, D., & Svoboda, K. (2010). Neural activity in barrel cortex underly- ing vibrissa-based object localization in mice. Neu- ron 67 (6), 1048–1061. [46] Roxin, A., Brunel, N., Hansel, D., Mongillo, G., & van Vreeswijk, C. (2011). On the distribution of firing rates in networks of cortical neurons. J. Neurosci. 31 (45), 16217–16226. [47] Buzsaki, G., & Mizuseki, K. (2014). The log-dynamic brain: how skewed distributions affect network opera- tions. Nat. Rev. Neurosci. 15, 264–278. [48] Song, S., Sjöström, P., Reigl, M., Nelson, S., & Chklovskii, D. (2005). Highly nonrandom features of synaptic connectivity in local cortical circuits. PLoS Biol. 3 (3), e68. [49] Lefort, S., Tomm, C., Sarria, J.-C. F., & Petersen, C. C. H. (2009). The excitatory neuronal network of the C2 barrel column in mouse primary somatosensory cortex. Neuron 61, 301–316. [50] Koulakov, A. A., Hromadka, T., & Zador, A. M. (2009). Correlated connectivity and the distribution of firing rates in the neocortex. J. Neurosci. 29 (12), 3685–3694. [51] Avermann, M., Tomm, C., Mateo, C., Gerstner, W., & Petersen, C. (2012). Microcircuits of excitatory and inhibitory neurons in layer 2/3 of mouse barrel cortex. J. Neurophysiol. 107 (11), 3116–3134. [52] Ikegaya, Y., Sasaki, T., Ishikawa, D., Honma, N., Tao, K., Takahashi, N., Minamisawa, G., Ujita, S., & Mat- suki, N. (2013). Interpyramid spike transmission stabi- lizes the sparseness of recurrent network activity. Cereb. Cortex 23 (2), 293–304. [53] Roxin, A. (2011). The role of degree distribution in shaping the dynamics in networks of sparsely connected spiking neurons. Front. Comput. Neurosci. 5 (8). [54] Kuhn, A., Aertsen, A., & Rotter, S. (2004). Neuronal integration of synaptic input in the fluctuation-driven regime. J. Neurosci. 24 (10), 2345–2356. [55] Stein, R. B., Gossen, E. R., & Jones, K. E. (2005). Neuronal variability: Noise or part of the signal? na- trevnsci 6, 389–397. [56] Marder, E., & Goaillard, J.-M. (2006). Variability, com- pensation and homeostasis in neuron and network func- tion. Nat. Rev. Neurosci. 7 (7), 563–574. [57] Van Vreeswijk, C., & Sompolinsky, H. (1998). Chaotic balanced state in a model of cortical circuits. Neural Comput. 10 (6), 1321–1371. [58] Tsodyks, M., Mitkov, I., & Sompolinsky, H. (1993). Pat- tern of synchrony in inhomogeneous networks of oscil- lators with pulse interactions. Phys. Rev. Lett. 71 (8). [59] Golomb, D., & Rinzel, J. (1993). Dynamics of globally coupled inhibitory neurons with heterogeneity. Phys. Rev. E 48 (6), 4810–4814. [60] Neltner, L., Hansel, D., Mato, G., & Meunier, C. (2000). Synchrony in heterogeneous networks of spiking neu- rons. Neural Comput. 12 (7), 1607–1641. [61] Denker, M., Timme, M., Diesmann, M., Wolf, F., & Geisel, T. (2004). Breaking synchrony by heterogeneity in complex networks. Phys. Rev. Lett. 92 (7), 074103– 1–074103–4. [62] Mejias, J. F., & Longtin, A. (2012). Optimal hetero- geneity for coding in spiking neural networks. Phys. Rev. Lett. 108. [63] Stocks, N. G. (2000). Suprathreshold stochastic reso- nance in multilevel threshold systems. Physical Review Letters 84 (11), 2310. [64] Shamir, M., & Sompolinsky, H. (2006). Implications of neuronal diversity on population coding. Neural Com- 19 put. 18 (8), 1951–1986. [65] Chelaru, M. I., & Dragoi, V. (2008). Efficient coding in heterogeneous neuronal populations. Proc. Natl. Acad. Sci. USA 105 (42). [66] Osborne, L. C., Palmer, S. E., Lisberger, S. G., & Bialek, W. (2008). The neural basis for combinato- rial coding in a cortical population response. J. Neu- rosci. 28 (50), 13522–13531. [67] Padmanabhan, K., & Urban, N. N. (2010). Intrinsic biophysical diversity decorrelates neuronal firing while increasing information content. Nat. Neurosci. 13 (10), 1276–1282. [68] Marsat, G., & Maler, L. (2010). Neural heterogeneity and efficient population codes for communication sig- nals. J. Neurophysiol. 104 (5), 2543–2555. [69] Holmstrom, L. A., Eeuwes, L. B., Roberts, P. D., & Portfors, C. V. (2010). Efficient encoding of vocaliza- tions in the auditory midbrain. J. Neurosci. 30 (3), 802– 819. [70] Yim, M. Y., Aertsen, A., & Rotter, S. (2013). Impact of intrinsic biophysical diversity on the activity of spiking neurons. Phys. Rev. E 87. [71] Lengler, J., Jug, F., & Steger, A. (2013). Reliable neu- ronal systems: The importance of heterogeneity. PLoS One. [72] Mejias, J. F., & Longtin, A. (2014). Differential ef- fects of excitatory and inhibitory heterogeneity on the gain and asynchronous state of sparse cortical networks. Frontiers in computational neuroscience 8. [73] Tripathy, S. J., Padmanabhan, K., Gerkin, R. C., & Urban, N. N. (2013). Intermediate intrinsic diversity enhances neural population coding. Proc. Natl. Acad. Sci. USA 110 (20), 8248–8253. [74] Bernacchia, A., & Wang, X.-J. (2013). Decorrelation by recurrent inhibition in heterogeneous neural circuits. Neural Comput. 25 (7), 1732–1767. [75] Schemmel, J., Grübl, A., Meier, K., & Müller, E. (2006). Implementing synaptic plasticity in a VLSI spiking neu- ral network model. In Proceedings of the 2006 Interna- tional Joint Conference on Neural Networks (IJCNN), Vancouver, pp. 1–6. IEEE Press. [76] Badoni, D., Giulioni, M., Dante, V., & Del Giudice, P. (2006). An aVLSI recurrent network of spiking neurons with reconfigurable and plastic synapses. In Proceedings of the 2006 International Symposium on Circuits and Systems (ISCAS), Island of Kos, pp. 4. IEEE Press. [77] Hafliger, P. (2007). Adaptive WTA with an analog VLSI neuromorphic learning chip. IEEE Trans. Neu- ral Netw. 18 (2), 551–572. [78] Vogelstein, R., Mallik, U., Vogelstein, J., & Cauwen- berghs, G. (2007). Dynamically reconfigurable sili- con array of spiking neurons with conductance-based synapses. IEEE Trans. Neural Netw. 18 (1), 253–265. [79] Indiveri, G., Chicca, E., & Douglas, R. (2009). Artifi- cial cognitive systems: From VLSI networks of spiking neurons to neuromorphic cognition. Cognitive Compu- tation 1 (2), 119–127. [80] Serrano-Gotarredona, R., Oster, M., Lichtsteiner, P., Linares-Barranco, A., Paz-Vicente, R., Gomez- Rodriguez, F., Camunas-Mesa, L., Berner, R., Rivas- Perez, M., Delbruck, T., Liu, S.-C., Douglas, R., Hafliger, P., Jimenez-Moreno, G., Ballcels, A., Serrano- Gotarredona, T., Acosta-Jimenez, A., & Linares- Barranco, B. (2009). CAVIAR: A 45k neuron, 5M synapse, 12G connects/s AER hardware sensory - pro- cessing - learning - actuating system for high-speed vi- sual object recognition and tracking. IEEE Trans. Neu- ral Netw. 20 (9), 1417–1438. [81] Brink, S., Nease, S., Hasler, P., Ramakrishnan, S., Wun- derlich, R., Basu, A., & Degnan, B. (2013). A learning- enabled neuron array IC based upon transistor channel models of biological phenomena. IEEE Trans. Biomed. Circuits Syst. 7 (1), 71–81. [82] Benjamin, B., Gao, P., McQuinn, E., Choudhary, S., Chandrasekaran, A., Bussat, J., Alvarez-Icaza, R., Arthur, J., Merolla, P., & Boahen, K. (2014). Neurogrid: A mixed-analog-digital multichip system for large-scale neural simulations. Proc. IEEE 102 (5), 699–716. [83] Renaud, S., Tomas, J., Lewis, N., Bornat, Y., Daouzli, A., Rudolph, M., Destexhe, A., & Saïghi, S. (2010). PAX: A mixed hardware/software simulation platform for spiking neural networks. Neural Networks 23 (7), 905–916. [84] Yu, T., & Cauwenberghs, G. (2010). Analog VLSI biophysical neurons and synapses with programmable membrane channel kinetics. IEEE Trans. Biomed. Cir- cuits Syst. 4 (3), 139–148. [85] Davison, A., Brüderle, D., Eppler, J. M., Kremkow, J., Muller, E., Pecevski, D., Perrinet, L., & Yger, P. (2009). PyNN: a common interface for neuronal network simu- lators. Front. Neuroinformatics 2 (11). [86] Brüderle, D., Müller, E., Davison, A., Muller, E., Schemmel, J., & Meier, K. (2009). Establishing a novel modeling tool: A Python-based interface for a neuromorphic hardware system. Front. Neuroinformat- ics 3 (17). [87] Kaplan, B., Brüderle, D., Schemmel, J., & Meier, K. (2009). High-conductance states on a neuromorphic hardware system. In Proceedings of the 2009 Interna- tional Joint Conference on Neural Networks (IJCNN), Atlanta, pp. 1524–1530. IEEE Press. [88] Bill, J., Schuch, K., Brüderle, D., Schemmel, J., Maass, W., & Meier, K. (2010). Compensating inhomogeneities of neuromorphic VLSI devices via short-term synaptic plasticity. Front. Comput. Neurosci. 4 (129). [89] Brüderle, D., Bill, J., Kaplan, B., Kremkow, J., Meier, K., Müller, E., & Schemmel, J. (2010). Simulator- like exploration of cortical network architectures with a mixed-signal VLSI system. In Proceedings of the 2010 International Symposium on Circuits and Systems (IS- CAS), Paris, France. IEEE Press. [90] Mainen, Z. F., & Sejnowski, T. J. (1995). Reliability of spike timing in neocortical neurons. Science 268, 1503– 1506. [91] De la Rocha, J., Doiron, B., Shea-Brown, E., Kresimir, J., & Reyes, A. (2007). Correlation between neural spike trains increases with firing rate. Nature 448 (16), 802– 807. [92] Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. (2007). Numerical Recipes: The Art of Scientific Computing (3rd ed.). Cambridge University Press. [93] Prinz, A. A., Bucher, D., & Marder, E. (2004). Similar network activity from disparate circuit parameters. Nat. Neurosci. 7, 1345–1352. 20 [94] Pfeil, T., Scherzer, A.-C., Schemmel, J., & Meier, K. (2013). Neuromorphic learning towards nano second precision. In Neural Networks (IJCNN), The 2013 In- ternational Joint Conference on, pp. 1–5. IEEE Press. [95] Bair, W., Zohary, E., & Newsome, W. (2001). Corre- lated firing in Macaque visual area MT: time scales and relationship to behavior. J. Neurosci. 21 (5), 1676–1697. [96] Kohn, A., & Smith, M. A. (2005). Stimulus dependence of neuronal correlations in primary visual cortex of the Macaque. J. Neurosci. 25 (14), 3661–3673. [97] Moreno-Bote, R., & Parga, N. (2006). Auto- and cross- correlograms for the spike response of leaky integrate- and-fire neurons with slow synapses. Phys. Rev. Lett. 96, 028101. [98] Maass, W. (2014). Noise as a resource for computa- tion and learning in networks of spiking neurons. Proc. IEEE 102 (5), 860–880. [99] London, M., Roth, A., Beeren, L., Häusser, M., & Latham, P. E. (2010). Sensitivity to perturbations in vivo implies high noise and suggests rate coding in cor- tex. Nature 466 (1), 123–128. [100] Grytskyy, D., Tetzlaff, T., Diesmann, M., & Helias, M. (2013). A unified view on weakly correlated recurrent networks. Front. Comput. Neurosci. 7, 131. [101] Gewaltig, M.-O., & Diesmann, M. (2007). NEST (NEu- ral Simulation Tool). Scholarpedia 2 (4), 1430. [102] Davison, A., Brüderle, D., Eppler, J., Kremkow, J., Muller, E., Pecevski, D., Perrinet, L., & Yger, P. (2008). PyNN: a common interface for neuronal network simulators. Front. Neuroinformatics 2 (11). doi:10.3389/neuro.11.011.2008. [103] Helias, M., Tetzlaff, T., & Diesmann, M. (2013). Echoes in correlated neural systems. New J. Phys. 15, 023002. [104] Rosenbaum, R., & Josic, K. (2011). Mechanisms that modulate the transfer of spiking correlations. Neural Comput. 23 (5), 1261–1305. [105] Ribrault, C., Sekimoto, K., & Triller, A. (2011). From the stochasticity of molecular processes to the variabil- ity of synaptic transmission. natrevnsci 12, 375–387. [106] O'Leary, T., Williams, A. H., Franci, A., & Marder, E. (2014). Cell types, network homeostasis, and patho- logical compensation from a biologically plausible ion channel expression model. Neuron 82 (4), 809–821. [107] Giulioni, M., Camilleri, P., Mattia, M., Dante, V., Braun, J., & Del Giudice, P. (2012). Robust working memory in an asynchronously spiking neural network re- alized in neuromorphic VLSI. Front. Neurosci. 5 (149). [108] Schemmel, J., Brüderle, D., Grübl, A., Hock, M., Meier, K., & Millner, S. (2010). A wafer-scale neuromorphic hardware system for large-scale neural modeling. In Proceedings of the 2010 International Symposium on Circuits and Systems (ISCAS), Paris, pp. 1947–1950. IEEE Press. [109] Nordlie, E., Gewaltig, M.-O., & Plesser, H. E. (2009). Towards reproducible descriptions of neuronal network models. PLoS Comput. Biol. 5 (8), e1000456. [110] Sharp, T., Galluppi, F., Rast, A., & Furber, S. (2012). Power-efficient simulation of detailed cortical microcir- cuits on SpiNNaker. J. Neurosci. Methods 210 (1), 110– 118. Supplements Supplements 1: Power consumption i The Spikey system consumes approximately 6 W of power, and the chip itself less than 0.6 W. On the chip most In power is consumed by digital communication infrastructure, which is not part of the neuromorphic network. the following, we estimate the power consumption for a single synaptic event using the data set partly shown in Figure 5a. This emulation lasts T = 10 s in biological time and generates approximately 45 · 103 spikes. Considering the acceleration of the hardware network (104) and the synapse count per neuron (K = 15), the system generates 7 · 108 synaptic events per second in hardware time. If we consider the total power consumption of the Spikey chip, the upper bound of energy consumed by each synaptic transmission will be approximately 1 nJ. Because these measurements include the communication infrastructure and other support electronics to observe spike times and membrane traces, the real energy consumption for synaptic transmissions is estimated to be approximately ten times smaller. Network simulations on conventional supercomputers a far less energy efficient and consume tens of µJ for each synaptic transmission [110]. Supplements 2: Modification of the bisection method In each iteration of the bisection method that is used to calibrate the leak conductances of hardware neurons (Section II E), we evaluated the firing rate for each neuron by the median over L = 25 identical trials. However, if this measure is compared between consecutive identical iterations, temporal noise on time scales longer than the duration of one iteration may still lead to variability. In the original bisection method, the interval of possible solutions is halved after each iteration step [92]. To improve the convergence of this method in the context of our calibration we expanded the halved interval by 20% at both ends after each iteration. This prevents the algorithm to get stuck in an interval wrongly chosen by random fluctuations of the firing rates. Supplements 3: Effective weights We quantify the effect of a single spike of neuron j on the firing rate of a postsynaptic neuron i by the effective weight wij of the connection i ← j. Assuming that the activity of neuron i does not affect the activity of neuron j (i.e., the RAND case), we define wij as the cross-correlation between the spike trains sj(t) and si(t + τ ), where, due to causality, τ is positive. Then, we average the effective weight over the emulated network time T , and subtract the baseline determined by the average correlation for negative τ: (cid:104)sj(t)si(t + τ )(cid:105)tdτ − 1 τmin −τmin (cid:104)sj(t)si(t + τ )(cid:105)tdτ . (S1) (cid:90) τmax 1 wij = τmax 0 (cid:90) 0 SUP. FIG. 1. Distribution of effective weights w (a; log scale) and population-averaged effective weights ¯w (b) for different levels of heterogeneity, averaged over M = 100 network realizations. Symbols and error bars denote mean and one standard deviation, respectively. 0.00.20.40.60.81.0Heterogeneity a0.200.150.100.050.00Effective weight wi (s−1)0.00.20.40.60.81.0Heterogeneity a0.0500.0450.0400.0350.030Avg. effective weight ¯w (s−1)10-410-310-2Rel. frequencyab ii SUP. FIG. 2. Distribution of average incoming effective weights wi over the rates of the postsynaptic neurons i, for three different levels of heterogeneity (a = 0, a = 0.625 and a = 1), averaged over M = 100 network realizations. The average incoming effective weights are obtained by summing over all presynaptic partners ( wi =(cid:80) j wij). Here, we chose τmax = 50 ms and τmin = 50 ms. τmax was determined by measuring the average duration in which a spike from neuron j has an influence on neuron i (data not shown). τmin was then chosen symmetrically. The mass of the effective weights density shifts towards less negative effective weights for increasing heterogeneity (Supplements Figure 1a). We obtain ¯w by averaging over all possible connections: (cid:88) i,j ¯w = 1 N K wij . (S2) For increasing heterogeneity the absolute value of the average effective weight ¯w decreases (Supplements Figure 1b). This can be explained by the dependence of the effective weight on the firing rate of the postsynaptic neuron (Sup- plements Figure 2). In the regime of small rates (< 10 s−1), incoming spikes hardly affect the neuron's firing and the effective weight is small. Similarly, in the case for large rates (> 40 s−1). Neurons with intermediate firing rates are sensitive to input and hence have a more negative effective weight. Heterogeneity increases the number of neurons with small and large firing rates, and hence the absolute value of the average effective weight decreases, which in turn weakens the effective negative feedback of the network. Supplements 4: Simulations with software We validate our results by comparing them to simulations with software (NEST [101], PyNN [102]). In these simulations we modulated the degree of heterogeneity by distributing the firing thresholds of all neurons according to a normal distribution with mean Θ and standard deviation σΘ. Details about the network, neuron and synapse models and their parameters can be found in Supplements Table I and II, respectively. The results are qualitatively similar to network emulations on the Spikey chip (compare Figure 8 to 9) and also hold for larger network sizes (Supplements Figure 3). In the FB case, input correlations and output correlations increase with network heterogeneity. In contrast to hardware emulations, input correlations stay approximately constant in the RAND case. Output correlations strongly decrease with the standard deviation σΘ. Here, heterogeneity only affects the output spike times, and not the integrative properties of the neurons (see also Section III B and [70]). In addition, we compare the effect of heterogeneity in firing thresholds on correlations to that of heterogeneity in synaptic weights. Details about the network, neuron and synapse models and their parameters can be found in Supplements Table I and III, respectively. If we distribute synaptic weights, shared-input correlations decrease with the width of the weight distribution (Supplements Figure 4). Output correlations are also reduced, however, only proportionally to the reduction of input correlations (compare insets in Supplements Figure 4). While output correlations are overall smaller than input correlations due to the non-linearity in spike generation, we do not observe a boost of this decrease for large heterogeneities. Overall, the dynamics of the recurrent system is more sensitive to heterogeneities in firing thresholds than in synaptic weights (compare scale of abscissas of Figure 9 to Supplements Figure 4). To investigate the generalization of our results to mixed excitatory-inhibitory networks, correlations were measured 04080120160Rate of neuron i (s−1)0.20000.12670.05330.0200Effective weight wi (s−1)a=1.004080120160a=0.62504080120160a=0.010-410-310-2Rel. frequency iii SUP. FIG. 3. Like Figure 9, but for N = 4800 neurons and K = 375 inputs for each neuron. Note the different scales of the ordinate in Figure 9 and (b). Symbols and error bars denote mean and standard deviation, respectively, across M = 20 network realizations (error bars are partly covered by markers). SUP. FIG. 4. Dependence of population-averaged input correlations (a), and spike-train correlations (b) on the width of the weight distribution, for the intact network (FB, dark gray circles) and the RAND (light gray circles) case. Symbols and error bars denote mean and standard deviation, respectively, across M = 30 network realizations (error bars are partly covered by markers). The gray line in (a) depicts shared-input correlations in a homogeneous network (Equation 2). The insets show correlations normalized to unity at σJ = 0, with the same abscissa as in the main plot. Note that in simulations the FBreplay is identical to the FB case, and is hence not shown. Networks simulated with NEST [101] and PyNN [102]. in a network of N = 192 neurons consisting of half excitatory and half inhibitory neurons. Details about the network, neuron and synapse models and their parameters can be found in Supplements Table IV and V, respectively. As described above, we distribute the firing thresholds of all neurons according to a normal distribution. The results are consistent with those obtained from purely inhibitory networks on hardware demonstrating the generality of our findings (compare Figure 8 to Supplements Figure 5). By modulating the level of heterogeneity separately for the excitatory or the inhibitory population, we observe that network dynamics are more sensitive to heterogeneities in the inhibitory than in the excitatory population (data not shown). We investigate the effect of temporal noise on correlations by dithering spikes in the FBreplay case before replaying them to the network (for network, neuron and synapse models see Supplements Table I and II). Each spike time t is replaced by a spike time t(cid:48) randomly drawn from a normal distribution with width ϑ: t(cid:48) ∼ N (t, ϑ). Even for small ϑ, correlations increase on the input as well as on the output side, demonstrating the sensitivity of correlations to perturbations in the feedback loop (see Supplements Figure 6 and [5]). This effect is also reflected in an increase of the power of the population activity at small frequencies (see Supplements Figure 7). These results suggest that on hardware temporal noise is responsible for the increase of correlations and population power in the FBreplay case (compare FBreplay to FB in Figure 5, 7 and 8). 0.000.020.040.060.080.100.120.14Rel. std. deviation of threshold potential σΘµΘ0.0000.0020.0040.0060.008Correlation κS0.000.020.040.060.080.100.120.14Rel. std. deviation of threshold potential σΘµΘ0.000.020.040.060.080.10Correlation κVFBRAND0.00020.00010.0000ab0.00.51.01.52.02.53.0Rel. std. deviation of weights σJµJ0.010.000.010.020.030.040.05Correlation κS0.00.51.01.52.02.53.0Rel. std. deviation of weights σJµJ0.000.020.040.060.080.10Correlation κVFBRAND0.51.00.51.0ab iv SUP. FIG. 5. Dependence of population-averaged input correlations (a), and spike-train correlations (b) on the width of the threshold distribution, for the intact network (FB, dark gray circles) and the RAND (light gray circles) case. The network consists of two populations, one excitatory and one inhibitory, of equal size. Symbols and error bars denote mean and standard deviation, respectively, across M = 30 network realizations (error bars are partly covered by markers). The gray line in (a) depicts shared-input correlations in a homogeneous network (Equation 2). The inset in (b) shows a magnified view of the spike-train correlations in the FB case. Note that in simulations the FBreplay is identical to the FB case, and is hence not shown. Networks are simulated with NEST [101] and PyNN [102]. SUP. FIG. 6. Dependence of population-averaged input correlations (a), and spike-train correlations (b) on the strength of temporal noise, for the FBreplay (black circles) and the RAND (light gray circles) case. Symbols and error bars denote mean and standard deviation, respectively, across M = 30 network realizations (error bars are partly covered by markers). Gray curve in (a) depicts shared-input correlations in a homogeneous network (Equation 2). Correlations in the FB case correspond to correlations in the FBreplay case for ϑ = 0 ms. Networks simulated with NEST [101] and PyNN [102]. Supplements 5: Linear model We investigate the consistency of our results with a linear rate model that allows us to numerically calculate the average correlations from a given connectivity matrix W. The model is defined as (according to, e.g., [100]) r(t) = (W(r + x) ∗ h)(t) . (S3) Here, r(t) denotes the rate of the individual neurons and x(t) a Gaussian white noise input that is independent for each neuron. The linear filter kernel h(t) depends on the details of the model, is not relevant in our calculation, and hence is not further specified here. Equation S3 can be transformed to Fourier domain, where the input and output 0.000.020.040.060.080.100.120.14Rel. std. deviation of threshold potential σΘµΘ0.010.000.010.020.030.040.05Correlation κS0.000.020.040.060.080.100.120.14Rel. std. deviation of threshold potential σΘµΘ0.000.020.040.060.080.10Correlation κVFBRANDEEEIII0.0100.0050.000ab0102030405060Dither of input spikes ϑ (ms)0.010.000.010.020.030.040.05Correlation κS0102030405060Dither of input spikes ϑ (ms)0.000.020.040.060.080.10Correlation κVRANDFBreplayab v SUP. FIG. 7. Power spectrum of the population activity of the networks in Supplements Figure 6 for the intact network (solid dark gray), different degrees of dither on input spikes (dashed, dashed dotted and dotted black) and the RAND case (solid light gray), averaged across M = 30 network realizations. spectral matrices can be expressed by (S4) (S5) with T (ω) = (1 − H(ω)W)−1 [100]. In the RAND case, the linear equation for the rate of the (unconnected) neurons reads CRR(ω) = T (ω)T (ω)† , RR(ω) = WCRR(ω)WT Cin , q(t) = (W(r + x) ∗ h)(t) , (S6) where r(t) has the same auto-correlations as r(t) but zero cross correlations, i.e., C R R = diag(CRR), since the randomization of spike times removes all spatio-temporal correlations. According to Tetzlaff et al. [5] spectral matrices in the RAND case are given by We calculate the population-averaged power- and cross-spectra from the full matrices: ¯AX(ω) = 1 N CXX,ii (cid:88) i 1 QQ(ω) =WC R RWT Cin CQQ(ω) =H(ω)2(Cin , QQ + ρ) . ¯CXX(ω) = N (N − 1) , CXX,ij . (cid:88) i(cid:54)=j (S7) (S8) (S9) (S10) Here, X ∈ {R, Q} denotes the FB and RAND case, respectively. The low frequency coherence is the cross-spectra normalized by the power spectra: κX (0) = ¯CXX(0) ¯AX(0) . (S11) Note that in the linear model we are actually taking the zero frequency coherence. As in the spiking model, we consider a sparse network, i.e., we randomly choose for each neuron i ∈ [1, N ] an identical number of presynaptic partners (K = 15). In the linear model we do not consider a distribution of non-zero effective weights. Instead, each realized connection is assigned the same weight value −w. To mimic the effect of calibration we vary the absolute value of the effective weight by scaling the weights of the non-zero connections with 100101102Frequency f (Hz)10-1100101102Power spectrum ¯A (s−1)FBDither (2ms)Dither (10ms)Dither (40ms)RAND vi SUP. FIG. 8. Dependence of population-averaged input correlations (a), and output correlations (b) on the absolute magnitude of the effective weights ("heterogeneity" a, see text), for the intact network (FB, dark gray diamonds) and the RAND (light gray circles) case. Gray curve in (a) depicts shared-input correlations in a homogeneous network (Equation 2). The inset in (b) shows a magnified view of the spike-train correlations in the FB case (dark gray diamonds). Note that in the linear model the FBreplay is identical to the FB case, and is hence not shown. SUP. FIG. 9. Distributions of effective weights in a sparse network. Different values for the mean µw and standard deviation σw of the distribution of non-zero weights in (a) and (b), respectively, result in different values for the mean µW and standard deviation σW of the effective weight matrix. a sigmoidal function of a ∈ [0, 1]: w = 1 1 + e10×(a−0.5) w . (S12) This procedure changes the variance of the weight matrix [74] and hence a is denoted the heterogeneity of the network. More homogeneous (heterogeneous) networks have larger (smaller) effective weights and hence stronger (weaker) feedback. We obtain qualitatively similar results as we observe on the Spikey chip (compare Supplements Figure 8 to Figure 8). Output correlations in the RAND case decrease, while both input and output correlations in the FB case increase with network heterogeneity, i.e., with the variance of the effective weight matrix. Note that, in this simplified model, input correlations in the RAND case are not affected by the level of "heterogeneity" because the non-zero weights are homogeneous, i.e. have zero variance, irrespectively of a (cf. Equation 3). In Supplements Figure 9 we illustrate, how in a sparse network the variance of the weight matrix can increase, although the distribution of non-zero weights becomes narrower. The standard deviation σw of a distribution of non-zero weights with mean µw is (in this example) smaller than the standard deviation σW of the full effective weight matrix, due to the sparseness of the matrix (Supplements Figure 9a; here, we chose  = 0.8). If we, at the same time, increase the mean µw and decrease the standard deviation σw of non-zero weights, the standard deviation of the weight matrix σW can increase significantly (Supplements Figure 9). While the distribution of effective weights 0.00.20.40.60.81.0Heterogeneity a0.010.000.010.020.030.040.05Output correlations0.00.20.40.60.81.0Heterogeneity a0.000.020.040.060.08Input correlationsFBRAND-0.004-0.0020.000ab0.00.20.40.60.8Effective weight (a.u.)10-310-210-1100Relative countµw=0.37, σw=0.04,µW=0.21, σW=0.190.00.20.40.60.8Effective weight (a.u.)µw=0.67, σw=0.01,µW=0.37, σW=0.33ab vii SUP. FIG. 10. CVISI distributions (0.54, 0.53, 0.5) are marked with dashed lines. Percentage of dead neurons: 0.00%, 0.02%, 0.08%. (a–c) Like Figure 6d–f, but for the RAND case. The mean of firing rate (24.0 s−1, 25.2 s−1, 27.8 s−1) and Firing-rate dependence of pairwise low-frequency spike-train coherence in the uncalibrated (a) and fully SUP. FIG. 11. calibrated system (b). Each bin represents the low-frequency spike-train coherence κij = , averaged across all neuron pairs {i, j} with firing rates in the interval [ri, ri + ∆r) and [rj, rj + ∆r). Here, Re(Cij(f )) denotes the real part of the complex cross-spectrum at frequency f. fmin = 0.1 Hz, fmax = 20 Hz, ∆r = 5 s−1. Data averaged across M = 100 network realizations. Re(Cij (f )) Ai(f )Aj (f ) fmax−fmin 1 (cid:82) fmax fmin √ is broadened, the mean is decreased, which has a greater impact on the size of correlations. This observation could explain the decrease of correlations with increased calibration of the neuromorphic chip. Supplements 6: Firing statistics in the RAND case The firing rate distributions in the RAND case are similar to those in the FB scenario (compare Supplements Figure 10 to Figure 6d–f). As in the FB case they become broader for increasing heterogeneity. Due to a larger variance of the membrane potentials, inactive neurons in the FB case have a higher probability to fire in the RAND case (compare gray bars in horizontal histograms of Supplements Figure 10a–c and Figure 6d–f). Neurons with firing rates larger than the average firing rate are strongly driven by constant current influx, and hence show similar firing rates than in the FB scenario. The irregularity of firing increases for the RAND compared to the FB case. This can also be traced back to an increased variance of the membrane potential for the RAND case (data not shown). Supplements 7: Correlation matrices In addition to the population-averaged measures from the main manuscript, we calculated the pairwise correlations for each pair i, j of neurons with i (cid:54)= j ∈ [1, N ], and ordered these by the time-averaged rate of the corresponding 04080120Rate r (s−1)0.00.51.01.52.02.5CVISI0408012004080120abc0255075100125150Rate (s−1)a=10255075100125150Rate (s−1)0255075100125150Rate (s−1)a=00.0100.0080.0060.0040.0020.0000.0020.0040.0060.0080.010Correlationab viii SUP. FIG. 12. Like Figure 4 and 8 in (A) and (B), respectively, but for two additional chips of version 5. neurons (Supplements Figure 11). This reveals a dependence of the pairwise correlation on the firing rate of the respective neurons. If both neurons fire at low rate (here < 5 s−1), correlations will be close to zero similar to the results in [91]. For high rates (here > 25 s−1) we find mostly positive correlations. However, for neurons firing at intermediate rates, the activity of the pair is often anti-correlated, and can hence effectively suppress (positive) shared-input correlations. After calibration, the amount of neurons firing at intermediate rates and showing negative correlations increases (Supplements Figure 11b, Figure 6). Shared-input correlations are hence suppressed by more neurons leading to smaller input correlations than in the uncalibrated case (Figure 8). Supplements 8: Results for different Spikey chips We have performed the same experiments as described in the main text using two additional Spikey chips of the current version (5). Different chips show different realizations of fixed-pattern noise, and hence calibration was repeated for each chip separately (Figure 4, Supplements Figure 12A). For all chips we find qualitatively the same results as described above: in the FB case, free membrane potentials (and spike trains) are decorrelated by inhibitory feedback and correlations increase with an increasing level of heterogeneity (Figure 8, Supplements Figure 12B). This also applies to the results of identical experiments on three previous-generation chips (version 4). These old chips show more pronounced heterogeneity in their parameters. In addition, we observe larger quantitative differences between old chips for the uncalibrated case, which is likely to be caused by different extents of intrinsic fixed-pattern noise. Unfortunately, the old systems show an artifact affecting the population power at around 100 Hz (Supplements Figure 13), consistently across all chips of identical revision. We can exclude that this effect can be traced back to network effects (it occurs across the FB, FBreplay and RAND cases) or to single-cell power spectra (data not shown). We hence have to consider it an artificial spurious synchronization. As the artifact lies outside the range of frequencies that are relevant for our measures of correlation it does not affect the main results and experiments on the old chips qualitatively confirm our results in the main manuscript (compare Supplements Figure 14 to Figure 4, 8 and Supplements Figure 12). Supplements 9: Reproducibility of networks with intact feedback We measured the standard deviation of free membrane potential and spike train correlations over several trials for a single network realization (Supplements Figure 15). This standard deviation is smaller than the standard deviation we observe over different network realizations (compare to Figure 8 and Supplements Figure 12), which indicates that (A)Calibration050100150Rate r(s−1)050100150200Neuron counta=1a=00.00.51.0Heterogeneity a010203040∆r (s−1)ab(B)Correlation0.00.20.40.60.81.0Heterogeneity a0.010.000.010.020.030.040.05Correlation κSSpike trains0.00.20.40.60.81.0Heterogeneity a0.000.020.040.060.08Correlation κVFree membrane potentialsRANDFBreplayFB0.00380.00320.0050.0100.0150.020ab(A)Calibration050100150Rate r(s−1)050100150200Neuron counta=1a=00.00.51.0Heterogeneity a010203040∆r (s−1)ab(B)Correlation0.00.20.40.60.81.0Heterogeneity a0.010.000.010.020.030.040.05Correlation κSSpike trains0.00.20.40.60.81.0Heterogeneity a0.000.020.040.060.08Correlation κVFree membrane potentialsRANDFBreplayFB0.00380.00320.0050.0100.0150.020ab ix SUP. FIG. 13. Spectra of trial-averaged population power ¯A(f ) and trial-averaged single-cell power A(f ) for chip 666 (version 4, left) and chip 508 (version 5, right) illustrating the synchronization artifact in the 100 Hz range for the intact network (FB, dark gray), the FBreplay (black) and the RAND (light gray) case. SUP. FIG. 14. Like Figure 4 and 8 in (A) and (B), respectively, but for three chips of version 4. 100101102Frequency f (Hz)10-1100101102Power spectrum ¯A(s−1)10010110210-1100101102A(s−1)100101102Frequency f (Hz)10-1100101102Power spectrum ¯A(s−1)10010110210-1100101102A(s−1)(A)Calibration050100150Rate r(s−1)050100150Neuron counta=1a=00.00.51.0Heterogeneity a010203040∆r (s−1)ab(B)Correlation0.00.20.40.60.81.0Heterogeneity a0.010.000.010.020.030.040.050.06Correlation κSSpike trains0.00.20.40.60.81.0Heterogeneity a0.000.020.040.060.08Correlation κVFree membrane potentialsRANDFBreplayFB0.0040.003ab(A)Calibration050100150Rate r(s−1)050100150Neuron counta=1a=00.00.51.0Heterogeneity a010203040∆r (s−1)ab(B)Correlation0.00.20.40.60.81.0Heterogeneity a0.010.000.010.020.030.040.050.06Correlation κSSpike trains0.00.20.40.60.81.0Heterogeneity a0.000.020.040.060.08Correlation κVFree membrane potentialsRANDFBreplayFB0.0040.003ab(A)Calibration050100150Rate r(s−1)050100150Neuron counta=1a=00.00.51.0Heterogeneity a010203040∆r (s−1)ab(B)Correlation0.00.20.40.60.81.0Heterogeneity a0.010.000.010.020.030.040.050.06Correlation κSSpike trains0.00.20.40.60.81.0Heterogeneity a0.000.020.040.060.08Correlation κVFree membrane potentialsRANDFBreplayFB0.0040.003ab x SUP. FIG. 15. Like Figure 8 and for the same chip, but for L = 20 trials of one network realization. the latter is mostly caused by the different realizations of the connectivity, not by trial-to-trial variability. Note that the variability between trials of networks with intact feedback is likely to be larger than between replays of network activity as shown in Figure 3, because network dynamics may be chaotic. The data shown in Supplements Figure 15 has to be interpreted with care, because the reproducibility of only a single network realization is considered. For different realizations, the mean and standard deviation will in general be different. Supplements 10: Extraction of effective PSP amplitudes, time constants and delays In order to provide an estimation for the order of magnitude of various neuron parameters, we fitted the postsynaptic potentials (PSPs) of current-based LIF neurons with exponential postsynaptic currents (PSCs) to the spike-triggered average (STA) of free membrane potentials obtained from hardware emulations. The differential equations describing the dynamics of membrane potentials Vj(t) and synaptic currents Ij(t) for this simplified model are given by τ eff m τ eff syn Vj(t) = − Vj(t) + RIj(t) Ij(t) = − Ij(t) + τ eff syn Jjksk(t − deff) . (S13) (S14) (cid:88) k syn the synaptic time constant and Jjk where R is the resistance of the membrane, τ eff the synaptic weight of the connection from neuron k to j. Here the superscript eff indicates that these are not the parameters realized on hardware, but of an abstract model that we fit to hardware measurements. Data were obtained from 100 network realizations of the RAND case (same dataset as for Figure 8). For each neuron we computed the STA of its free membrane potential individually for each spike train of its presynaptic partners. The PSP for the current based model described above is given by m the membrane time constant, τ eff (cid:16) V (t) =V eff max τ eff syn syn − τ eff τ eff m (cid:17) e−(t−deff)/τ eff syn − e−(t−deff)/τ eff m Θ(t − deff) , (S15) max, τ eff where we introduced V eff syn, τ eff parameters: V eff max = RJ to simplify the fitting procedure. For each pair of neurons i, j we obtain a set of The fitted curves (using the function scipy.optimize.curve_fit with default parameters from Python's scipy module v0.17.0) match fairly well to the empirical data to allow an estimation of the effective values of the neuron parameters governing the shape of the postsynaptic potentials (Supplements Figure 16a). However, note that the resulting parameter distributions do not precisely reflect the actual parameter distributions on the chip. m and deff. The distributions of neuron and synapse parameters change slightly with calibration. For the calibrated network the m and deff are larger than for the uncalibrated network (see Table II and Supplements max, τ eff syn, τ eff absolute values of V eff Figure 16b). 0.01.0Heterogeneity a-0.010.00.010.020.030.040.050.06Correlation κSSpike trainsFBFBreplayRAND0.01.0Heterogeneity a0.00.020.040.060.08Correlation κVFree mem. pot.ab xi SUP. FIG. 16. (a) Fits of theoretical PSPs (dashed; for model, see Equation S15) to STAs of free membrane potentials in the FB case (solid). (b) Distributions of effective parameters obtained from fitting. Black and gray lines correspond to the fully calibrated (a = 0) and uncalibrated (a = 1) network, respectively. Median, 25th and 75th percentile of all fitted parameters are reported in Table II. 50510152025303540Time (ms)4.03.53.02.52.01.51.00.50.00.5STA (mV)a6050403020100Synaptic weight RJ10-310-210-1Rel. count01020304050Membrane time constant τm0.000.020.040.060.080.100.120.140510152025Synaptic time constant τs0.000.050.100.150.200.25Rel. count01234567Delay d0.00.20.40.60.81.01.21.41.6b xii A Populations Topology Connectivity Neuron model Channel models Synapse model Plasticity External input Measurements Other B Name I Model summary One (inhibitory) - Random convergent connections (fixed in-degree) Leaky integrate-and-fire (LIF), fixed firing threshold, fixed absolute refractory time - Exponentially decaying currents, fixed delays - Resting potential higher than threshold (= constant current) (El > Θ) Spikes and membrane potentials No autapses, no multapses Populations Elements LIF neuron Size N C Source Target Pattern I I Connectivity Random convergent connect, in-degree K Neuron and synapse model Leaky integrate-and-fire, exponential currents Subthreshold dynamics (t (cid:54)∈ (t∗, t∗ + τref)): Reset and refractoriness (t ∈ (t∗, t∗ + τref)): dt v(t) = −gl(v(t) − El) + Isyn(t) Cm d v(t) = vreset dt Isyn(t) = −Isyn(t) +(cid:80) d τsyn i,k Jδ(t − tk i ) D Type Subthreshold dynamics Current dynamics Spiking E Spike trains Membrane potentials Here the sum over i runs over all presynaptic neurons and the sum over k over all spike times of the respective neuron i If v(t∗−) < Θ ∧ v(t∗+) ≥ Θ: emit spike with time stamp t∗ Measurements recorded from Ps neurons recorded from Pv neurons SUP. TABLE I. Description of the network model (according to [109]). xiii B Name N C Name K D Name Cm τref vreset El Θ gl D Name τsyn J d E Name Ps Pv Values {192, 4800} Values {15, 375} Values 0.2 nF 0.1 ms −80 mV −52 mV ∼ N (−62, [0, 8.8]) mV 10 nS Values 5 ms −0.254 nA 1.0 ms Values {192, 4800} {150, 150} Populations Description network size Connectivity Description number of presynaptic partners Neuron Description membrane capacitance refractory period reset potential resting potential firing threshold leak conductance Synapse Description synaptic time constant synaptic weight synaptic delay Measurements Description number of neurons spike trains are recorded from number of neurons membrane potentials are recorded from SUP. TABLE II. Parameter values for the network model described in Supplements Table I with distributed thresholds. xiv B Name N C Name K D Name Cm τref vreset El Θ gl D Name τsyn J d E Name Ps Pv Values 192 Values 15 Values 0.2 nF 0.1 ms −80 mV −52 mV −62 mV 10 nS Populations Description network size Connectivity Description number of presynaptic partners Neuron Description membrane capacitance refractory period reset potential resting potential firing threshold leak conductance Values 5 ms ∼ [N (0.254, [0.00254, 0.75])]+ 1.0 ms Synapse Description synaptic time constant synaptic weight, clipped to positive values synaptic delay Values 192 150 Measurements Description number of neurons spike trains are recorded from number of neurons membrane potentials are recorded from SUP. TABLE III. Parameter values for the network model described in Supplements Table I with distributed weights. xv A Populations Topology Connectivity Neuron model Channel models Synapse model Plasticity External input Measurements Other B Name E I Model summary Two (excitatory, inhibitory) - Random convergent connections (fixed in-degree) Leaky integrate-and-fire (LIF), fixed firing threshold, fixed absolute refractory time - Exponentially decaying currents, fixed delays - Resting potential higher than threshold (= constant current) (El > Θ) Spikes and membrane potentials No autapses, no multapses Populations Elements LIF neuron LIF neuron Size NE NI Connectivity Random convergent connect, in-degree KE, weight JE Random convergent connect, in-degree KE, weight JE Random convergent connect, in-degree KI, weight JI Random convergent connect, in-degree KI, weight JI Neuron and synapse model Leaky integrate-and-fire, exponential currents Subthreshold dynamics (t (cid:54)∈ (t∗, t∗ + τref)): Reset and refractoriness (t ∈ (t∗, t∗ + τref)): dt v(t) = −gl(v(t) − El) + Isyn(t) Cm d v(t) = vreset dt Isyn(t) = −Isyn(t) +(cid:80) d τsyn i,k Jδ(t − tk i ) C Source Target Pattern E E I I E I E I D Type Subthreshold dynamics Current dynamics Spiking E Spike trains Membrane potentials Here the sum over i runs over all presynaptic neurons and the sum over k over all spike times of the respective neuron i If v(t∗−) < Θ ∧ v(t∗+) ≥ Θ: emit spike with time stamp t∗ recorded from P E recorded from P E s v Measurements excitatory and P I excitatory and P I s v inhibitory neurons inhibitory neurons SUP. TABLE IV. Description of the network model consisting of an excitatory and an inhibitory population (according to [109]). B Name NE NI C Name KE KI D Name Cm τref vreset El Θ gl D Name τsyn JE JI d E Name P E s P I s P E v P I v Values 96 96 Values 7 8 Values 0.2 nF 0.1 ms −80 mV −52 mV −62 mV 10 nS Values 5 ms 0.0635 nA −0.254 nA 1.0 ms Values 96 96 150 150 xvi Populations Description size of the excitatory population size of the inhibitory population Connectivity Description number of excitatory presynaptic partners number of inhibitory presynaptic partners Neuron Description membrane capacitance refractory period reset potential resting potential firing threshold leak conductance Synapse Description synaptic time constant excitatory synaptic weight inhibitory synaptic weight synaptic delay Measurements Description number of excitatory neurons spike trains are recorded from number of inhibitory neurons spike trains are recorded from number of excitatory neurons membrane potentials are recorded from number of inhibitory neurons membrane potentials are recorded from SUP. TABLE V. Parameter values for the network model described in Supplements Table IV.
1705.08261
1
1705
2017-05-17T15:51:37
Warnings and Caveats in Brain Controllability
[ "q-bio.NC", "physics.bio-ph" ]
In this work we challenge the main conclusions of Gu et al work (Controllability of structural brain networks. Nature communications 6, 8414, doi:10.1038/ncomms9414, 2015) on brain controllability. Using the same methods and analyses on four datasets we find that the minimum set of nodes to control brain networks is always larger than one. We also find that the relationships between the average/modal controllability and weighted degrees also hold for randomized data and the there are not specific roles played by Resting State Networks in controlling the brain. In conclusion, we show that there is no evidence that topology plays specific and unique roles in the controllability of brain networks. Accordingly, Gu et al. interpretation of their results, in particular in terms of translational applications (e.g. using single node controllability properties to define target region(s) for neurostimulation) should be revisited. Though theoretically intriguing, our understanding of the relationship between controllability and structural brain network remains elusive.
q-bio.NC
q-bio
Warnings and Caveats in Brain Controllability Chengyi Tu1,7, Rodrigo P. Rocha1,2,7, Maurizio Corbetta3,4,7, Sandro Zampieri5,7, Marzo Zorzi6,7,8 & S. Suweis1,7∗ 1Dipartimento di Fisica e Astronomia, 'G. Galilei' & INFN, Università di Padova, Padova, IT. 2Departamento de Física, Universidade Federal de Santa Catarina, 88040-900, Florianópolis-SC, Brazil. 3Dipartimento di Neuroscienze, Università di Padova, Padova, IT. 4Departments of Neurology, Radiology, Neuroscience, and Bioengineering, Washington University, School of Medicine, St. Louis, USA. 5Dipartimento di Ingegneria dell'informazione, Università di Padova, Padova, IT. 6Dipartimento di Psicologia Generale, Università di Padova, Padova, IT. 7Padua Neuroscience Center, Università di Padova, Padova, IT. 8IRCCS San Camillo Hospital Foundation, Venice, IT, There is large consensus that the complex, self-organizing structure of the human brain can be well described by the mathematical framework based on network theory. A recent article by Gu et al.1 proposed to characterize brain networks in terms of their "controllability", drawing on concepts and methods of control theory. The analysis of controllability has the potential to unveil how specific nodes and/or sets of nodes control the dynamics of the entire network1-4 and thus might provide insights on whether manipulating the local activity of specific nodes would fully or partially restore network functions after brain damage. Gu et al. applied the tools of control theory to quantify how the networks structure, defined by human connectome data (i.e., white matter pathways derived from diffusion tensor or diffusion weighted imaging), constrain or facilitate changes in brain state trajectories. They proposed that single brain regions that are densely connected facilitate the brain to easily reach many of its cognitive states (i.e. high average controllability), while weakly connected nodes promote the movement of the brain to difficult-to-reach states (i.e. high modal controllability). In other words, they claim that brain activity is controllable by a single node, and 1 topology of brain networks provides an explanation for the types of control roles that different regions play in the brain. Here we present new results based on the analysis of four dataset and numerical simulations that challenge their main conclusion and undermine their use of the controllability framework. In order to study single node controllability (in Kalman sensu), the authors first normalize the matrix describing the brain structural connectivity, and then calculate the controllability Gramian from each node (see Methods). Their first finding is that the brain can be theoretically controllable by a single region/node, i.e. the smallest (in absolute value) of the eigenvalues of the controllability Gramian 𝜆!"# (see Methods) from each brain region as a control node is greater than zero. Then they measure the average controllability and the modal controllability from each node (see Methods), and find that average controllability is strongly correlated with the weighted degree, and the modal controllability is strongly anti-correlated with the weighted degree. Finally they evaluate the control contribution of different brain regions associated with known cognitive process (i.e. resting state networks (RSNs)) finding that different RSNs have different control roles. To investigate further their results, we apply the controllability framework on four empirical open access datasets of large-scale structural brain connectivity from different studies5–8. In contrast to Gu et al.1, we observe that for all the analyzed brain networks 𝜆!"# is not greater than zero (we note that already in their region as a control node is 𝜆!"# = 2.5×10!!"± 4.8×10!!" and thus it is not significantly greater than work, also Gu et al.1 have found that the smallest eigenvalue of their controllability Gramian from each brain zero). Indeed, according to our data, the controllability Gramian from each node is not invertible. Therefore, it is not possible to assess the global controllability of the brain performed by a single region/node. We highlight that the structural properties of the DTI/DSI network data (symmetric with no self-loops) crucially affect the controllability of the system. Moreover, it is important also to quantify the energy needed to control the system from a specific region. In fact, although Gu et al.1 have found that on average 𝜆!"# ~10!!", in order to control such a system from a single region (node) would need an energy ε≈𝜆!"#!!~10!", in practice the system is not 2 controllable. Our results are in accordance with prior theoretical results9 showing that for symmetric networks (like DTI/DSI brain connectivity data) the energy ε needed to control from a single node the system grows exponentially with the network size (i.e. to practically control the system (ε not too large) a fixed fraction of the networks nodes is necessary9). Next, we study the role of nodes topological properties in controlling the brain network. Inspired by Gu and collaborators claim on the importance of nodes centrality, we first rank all networks nodes according to five distinct centrality measures. Then we build two ranking lists, following respectively the ascending and descending order of centralities magnitude. Then we generate a list of nodes in random order, irrespective of the nodes properties. For each of these three nodes sequences, we look for the minimum subsets of nodes needed to control the brain network. We have repeated this experiment for all brain connectivity datasets as well as for synthetic brain networks generated with a desired network topology (see Methods and Table 1). If topology is important we expect that the minimum number of control nodes will depend on the specific nodes properties (e.g. the most or the least central nodes). Our results are presented in Table 1. It shows the size of the minimum subset of control nodes with respect to five different centralities measures for the APOE-4 data5 (𝑁=110 nodes), random Barabasi-Albert (BA) scale free networks10 and uniform networks10, both with the same number of nodes, edges, and edge weight distribution. No matter which centrality measure is adopted and how we generated the node sequences, the size of the minimum subset is always greater than one and similar in all three cases (see Table 1). This result strongly suggests that the topological structure of the DTI/DSI connectivity networks do not play an important role in brain controllability. We thus want to better understand the relationships between average/modal controllability and weighted degree and test if it is due to the specific brain network topology, or rather if it holds also for random networks obtained by randomizing the structure of real brain data, indeed, even if the controllability Gramian is not invertible, the average/modal controllability can still be calculated. Following Gu et al. (Fig. 2 b and d in1), we plot the average of the ranks for weighted degrees versus average of the ranks of the average controllability in Fig. 1a, and the average of ranks for weighted degrees versus the average of ranks of the modal controllability in Fig. 1c, (for details on average 3 rank plots see Gu et al.'s methods section) of our four dataset. As Fig. 1a and 1c show, we find the same relation between controllability and network centrality reported by Gu et al. We then randomize the structural connectivity data and calculate average/modal controllability as a function of nodes properties in the random networks. Surprisingly, as shown in Fig. 1b and 1d, we found exactly the same relationship as before. We stress that the relationships in average controllability and modal controllability between real and randomized networks are not an artifact of the rank-rank plots. In fact, controllability measures between real and randomized networks are not distinguishable also by examining unranked magnitudes. Finally we want to test if different RSNs are involved in different modes of controllability. To do that, we have built RSNs for the Hagmann dataset (based on RSNs template in the literature11). Following Gu et al. procedure, we consider the thirty nodes with highest average or modal controllability, and then calculate a vector cdata, where its j-th component gives the percentage of the nodes that belongs to the j-th RSN. We then repeat the same analysis for a proper random null model10,12, obtaining crand from random networks with the same degree distribution of the original dataset, but with no spatial correlations among regions (as observed in real RSNs). We find that both real and randomized data display very similar results on the modes of controllability of the different RSNs; indeed the Pearson correlation between cdata and crand is ρ =0.7 (P-value ≈ 0.05). In other words we find that there are not specific roles played by RSNs in controlling the brain. In summary, our results challenge the main conclusions of Gu et al work on brain controllability. Using the same methods and analyses on four datasets we find that the minimum set of nodes to control brain networks is always larger than one. We also find that the relationships between the average/modal controllability and weighted degrees also hold for randomized data and the there are not specific roles played by RSNs in controlling the brain. In conclusion, we show that there is no evidence that topology plays specific and unique roles in the controllability of brain networks. Accordingly, Gu et al. interpretation of their results1, 13–15, in particular in terms of translational applications (e.g. using single node controllability properties to define target region(s) for neurostimulation) should be revisited. Though theoretically intriguing, our understanding of the relationship between controllability and structural brain network remains elusive. 4 y t i l i b a l l o r t n o c e g a r e v a f o k n a r f o e g a r e v a y t i l i b a l l o r t n o c l a d o m f o k n a r f o e g a r e v a 250 a 200 150 100 50 0 250 200 150 100 50 0 c 0 0 y t i l i b a l l o r t n o c e g a r e v a f o k n a r f o e g a r e v a y t i l i b a l l o r t n o c l a d o m f o k n a r f o e g a r e v a 250 b 200 150 100 50 0 0 250 d 200 150 100 50 50 0 0 50 Rockland APOE-4 Autism Hagmann 50 100 150 200 250 average of rank of weighted degree Rockland APOE-4 Autism Hagmann 50 100 150 200 250 average of rank of weighted degree randomized Rockland randomized APOE-4 randomized Autism randomized Hagmann 250 200 average of rank of weighted degree 100 150 randomized Rockland randomized APOE-4 randomized Autism randomized Hagmann 100 200 average of rank of weighted degree 150 250 Figure 1: Comparing controllability measures between empirical data and randomized data. Scatter plot of the average of the ranks for weighted degrees versus average of the ranks of the average controllability for (a) the empirical data and (b) their randomized counterpart. Scatter plot of the average of ranks for weighted degrees versus the average of ranks of the modal controllability for (c) empirical data and (d) their randomized counterpart. APOE-4 data Low 31/0.28/86.16 29/0.26/80.60 30/0.27/83.34 29/0.26/80.60 30/0.27/83.38 High 30/0.27/83.37 31/0.28/86.15 31/0.28/86.15 30/0.27/83.37 30/0.27/83.37 BA network Low 28.04/0.25/77.95 28/0.25/77.84 28.04/0.25/77.95 28/0.25/77.84 27.04/0.25/77.94 High 28.04/0.25/77.91 28.16/0.26/78.25 27.8/0.25/77.25 27.88/0.25/77.47 28.04/0.25/77.93 Uniform network Low 28.16/0.26/78.27 27.68/0.25/76.94 28/0.25/77.83 28.16/0.26/78.27 27.46/0.25/76.33 High 27.8/0.25/77.27 27.96/0.25/77.71 27.84/0.25/77.38 27.8/0.25/77.27 27.66/0.25/76.88 Centrality measure Degree centrality Betweenness centrality Eigenvector centrality Pagerank centrality Random sequence 5 Table 1: The minimum number / fraction (with respect the size of the network) of nodes and its energy Trace{WK} to guarantee the controllability of the APOE-4 data and the corresponding BA and uniform random networks (of same size, connectivity and edge weight distribution) following the procedure described in the main text. For a given sequence, "High" is the rank from the largest to the smallest, while "Low" is the rank from the smallest to the largest. Hagmann, Autism and Rockland dataset have been also analyzed and we found qualitatively the same results, i.e. in order to control the network we need a fraction of nodes ranging from 16% to 25% of the whole brain, regardless of the centrality measures used to rank the nodes. Methods Neural dynamics model. Similarly to Gu et. al., we employed a simplified linear discrete-time neural dynamics model. This model is a noise-free variation of the Galán model2 and can be derived from the linearization of a general Wilson-Cowan system (for strengths and weakness of this model see Refs. 𝑥𝑡+∆𝑡 =𝐴𝑥𝑡 +𝐵!𝑢!𝑡 (2),(3),(12)). We then discretized the linearized dynamics and obtained the following equation: where 𝑥𝑡 is the state variable representing the neural activity of the brain regions, 𝐴= 1−𝛼∆𝑡𝐼+ 𝑐𝑀∆𝑡 is the 𝑁×𝑁 Jacobian matrix, 𝛼 is the inverse of the relaxation time, ∆𝑡 is the discretized time step, 𝐼 is the 𝑁×𝑁 identity matrix, 𝑀 is the symmetric and weighted DTI structural matrix and c is a normalization constant. Following the original work of Galan2, we fixed 𝛼=1.0,∆𝑡=0.2. The input matrix 𝐵! identifies the set of control points K in the brain, where 𝐾={𝑘!,···,𝑘!} and 𝐵!= 𝑒!!,···,𝑒!! , where 𝑒! denotes the i-th canonical vector of dimension 𝑁. The input 𝑢!:𝑅!!→𝑅! denotes the control Controllability Gramian. Given the neural dynamics, Eq. (1), we can set up the Lyapunov equation, 𝑊!− 𝐴𝑊!𝐴!=𝐵!𝐵!! . If A is stable, then this equation has a unique solution called Gramian: strategy. (1) (2) 6 𝑊!= 𝐴!𝐵!𝐵!! !!!! 𝐴! ! (3) adopted by Gu et. al. The results presented in Table 1 were employed using the averaged DTI matrix, i.e. The results presented in the main text were obtained as follows. In Figure 1, we applied the controllability total number of healthy subjects. We calculate the average and the modal controllability for each of the For the discretized dynamics (1) and 𝐴 symmetric, stability is ensured if all eigenvalues of 𝐴 lies in the interval 𝜆!∈(−1,1) con i=1,…,N. Therefore, in order to guarantee stability of our dynamical system we set the normalization constant to 𝑐=1/(1+𝜆!"# ), where 𝜆!"#=𝑀𝑎𝑥{𝜆!,𝑖=1,…,𝑁}. Following this procedure we constrain the eigenvalue spectrum of 𝐴 to be in the desired interval. framework in each of the individual DTI matrices within a given dataset, 𝑀!, where 𝑖 ranges from 1 to 𝑛, the corresponding Jacobian matrix 𝐴! and then we performed the average of the rank plots following the procedure 𝑀!"= 𝑀! 𝑛. Trace{WK}. As shown by Gu and collaborators1, the strong correlation between node degree and average evolutionary mode of a dynamical network. It is given by 𝜙!= where 𝑉=𝜈!"! is !!!! the eigenvector matrix and 𝜆1,···,𝜆𝑁 are the corresponding eigenvalues of 𝐴. and edge weight distribution of the averaged DTI matrix, 𝑀!"= 𝑀! 𝑛 for each dataset. We implemented these constraints by fitting the edge weights of 𝑀!" to a Pareto distribution10. The synthetic networks were Barabasi-Albert scale-free networks10 and the uniform networks10, using the same number of nodes, edges follows the expressions reported by Gu et. al.1. The average controllability can be approximate by Controllability measures. The average and the modal controllability measures employed in this study Synthetic random networks. We generated two type of synthetic DTI random networks, namely, the controllability can be understood analytically and it does not depend on the specific topological properties of the structural connectivity M. The modal controllability represents the ability of a node to control each 1−𝜆!𝐴 !𝜈!"! generated using standard routines available in the Mathematica software. Controllability measures displayed in Table 1 were averaged over 50 realizations of the synthetic random networks. 7 Betweenness centrality is a measure of centrality based on shortest paths and is given by Centralities measures. Degree centrality refers to the number of links a node has to other nodes10. 𝑛!" 𝑣 𝑛!" where 𝑛!" is the total number of shortest paths from node 𝑠 to node 𝑡 and 𝑛!"𝑣 is the number of those paths that pass through the node 𝑣10. Eigenvector centrality is based on the centralities of its neighbor nodes and can be expressed as 𝑐=!!𝐴!𝑐, where 𝐴 is the adjacency matrix and 𝜆 is the corresponding largest eigenvalue10. !!!!! PageRank centrality is a generalization of the eingevector centrality, weighting in a non-linear way the number and quality of links connected to the node10. DTI brain connectivity datasets. We applied the controllability framework on four empirical datasets of large scale structural (DTI or DSI) brain connectivity from different studies 5–8. All datasets are open access and were obtained from the Human Connectome Project4, namely, the APOE-4 dataset5 (𝑛=30 APOE-4 non-carrier and n=25 APOE-4 carrier individuals; gray matter parcellation into 𝑁=110 large scale regions); The Rockland dataset6 (𝑛=195 healthy subjects, 𝑁=188 large scale regions); The Hagmann dataset8 (average matrix corresponding to 𝑛=5 healthy subjects, 𝑁=66 cortical regions) and the Autism dataset7 (𝑛=94 healthy subjects, 𝑁=264 large scale regions). For more specific details on the data acquisition and Randomized data set. In Figure 1 we randomized the topological structure of the DTI matrix 𝑀 by randomly random with zero diagonal entries. In building random RSNs, we randomized the DTI matrix 𝑀 by randomly rewiring all its matrix elements, but keeping symmetry. The randomized DTI matrices are thus symmetric preprocessing we refer to the original studies. rewiring all its matrix elements, but fixed the degree sequence of the original dataset12. References. 1. Gu, S., et al. Controllability of structural brain networks. Nature communications 6, 8414, doi:10.1038/ncomms9414 (2015). 2. Gala´n, R. F. On how network architecture determines the dominant patterns of spontaneous neural activity. PloS one 3, e2148 (2008). 8 3. Honey, C. J., et al. Predicting human resting-state functional connectivity from structural connectivity. Proc. Natl Acad. Sci. USA 106, 20352040 (2009). 4. USC Multimodal Connectivity Database: http://umcd.humanconnectomeproject.org/umcd/default/browse 5. Brown, J. A., et al. Brain network local interconnectivity loss in aging APOE-4 allele carriers. Proceedings of the National Academy of Sciences 108, 20760-20765 (2011). 6. Brown, J. A., Rudie, J. D., Bandrowski, A., Van Horn, J. D. and Bookheimer, S. Y. The UCLA multimodal connectivity database: a web-based platform for brain connectivity matrix sharing and analysis. Front Neuroinform 6, 28, doi:10.3389/fninf.2012.00028 (2012). 7. Brown, Jesse A., et al. The UCLA multimodal connectivity database: a web-based platform for brain connectivity matrix sharing and analysis. Frontiers in neuroinformatics 6: 28 (2012). 8. Hagmann, P., et. al. Mapping the Structural Core of Human Cerebral Cortex PLoS Biol 6(7):e159 (2008). 9. Pasqualetti, F., Zampieri, S., Bullo, F. Controllability metrics, limitations and algorithms for complex networks. IEEE Transactions on Control of Network Systems, 1(1), 40-52 (2014) 10. Newman, M. E. J. Networks: An Introduction. (Oxford University Press Inc., New York 2010). 11. Sizemore, A., Giusti, C., Betzel, R. F., & Bassett, D. S. Closures and cavities in the human connectome. arXiv:1608.03520 (2016). 12. Molloy, Michael, and Bruce Reed. "A critical point for random graphs with a given degree sequence." Random structures & algorithms 6.2‐3 (1995): 161-180. 13. Muldoon, Sarah Feldt, et al. Stimulation-based control of dynamic brain networks. PLoS Comput Biol 12.9: e1005076 (2016). 14. Gu, Shi, et al. Optimal Trajectories of Brain State Transitions. NeuroImage (2017). 15. Betzel, Richard F., et al. "Optimally controlling the human connectome: the role of network topology." Scientific Reports 6 (2016). Authors Contributions: SS., S.Z, M.Z., M.C. designed the research, C.T., R.R., S.S. performed the research, R.R., C.T. analyzed the data. All authors contributed in writing the manuscript. 9
1912.06207
1
1912
2019-12-12T20:54:08
From deep learning to mechanistic understanding in neuroscience: the structure of retinal prediction
[ "q-bio.NC", "cs.LG", "physics.bio-ph" ]
Recently, deep feedforward neural networks have achieved considerable success in modeling biological sensory processing, in terms of reproducing the input-output map of sensory neurons. However, such models raise profound questions about the very nature of explanation in neuroscience. Are we simply replacing one complex system (a biological circuit) with another (a deep network), without understanding either? Moreover, beyond neural representations, are the deep network's computational mechanisms for generating neural responses the same as those in the brain? Without a systematic approach to extracting and understanding computational mechanisms from deep neural network models, it can be difficult both to assess the degree of utility of deep learning approaches in neuroscience, and to extract experimentally testable hypotheses from deep networks. We develop such a systematic approach by combining dimensionality reduction and modern attribution methods for determining the relative importance of interneurons for specific visual computations. We apply this approach to deep network models of the retina, revealing a conceptual understanding of how the retina acts as a predictive feature extractor that signals deviations from expectations for diverse spatiotemporal stimuli. For each stimulus, our extracted computational mechanisms are consistent with prior scientific literature, and in one case yields a new mechanistic hypothesis. Thus overall, this work not only yields insights into the computational mechanisms underlying the striking predictive capabilities of the retina, but also places the framework of deep networks as neuroscientific models on firmer theoretical foundations, by providing a new roadmap to go beyond comparing neural representations to extracting and understand computational mechanisms.
q-bio.NC
q-bio
From deep learning to mechanistic understanding in neuroscience: the structure of retinal prediction Hidenori Tanaka1,2, Aran Nayebi3, Niru Maheswaranathan3,5, Lane McIntosh3, Stephen A. Baccus4, and Surya Ganguli2,5 1Physics & Informatics Laboratories, NTT Research, Inc., East Palo Alto, CA, USA 2Department of Applied Physics, Stanford University, Stanford, CA, USA 3Neurosciences PhD Program, Stanford University, Stanford, CA, USA 4Department of Neurobiology, Stanford University, Stanford, CA, USA 5Google Brain, Google, Inc., Mountain View, CA, USA Abstract Recently, deep feedforward neural networks have achieved considerable success in modeling biological sensory processing, in terms of reproducing the input-output map of sensory neurons. However, such models raise profound questions about the very nature of explanation in neuroscience. Are we simply replacing one complex system (a biological circuit) with another (a deep network), without understanding either? Moreover, beyond neural representations, are the deep network's computa- tional mechanisms for generating neural responses the same as those in the brain? Without a systematic approach to extracting and understanding computational mechanisms from deep neural network models, it can be difficult both to assess the degree of utility of deep learning approaches in neuroscience, and to extract experi- mentally testable hypotheses from deep networks. We develop such a systematic approach by combining dimensionality reduction and modern attribution methods for determining the relative importance of interneurons for specific visual computa- tions. We apply this approach to deep network models of the retina, revealing a conceptual understanding of how the retina acts as a predictive feature extractor that signals deviations from expectations for diverse spatiotemporal stimuli. For each stimulus, our extracted computational mechanisms are consistent with prior scientific literature, and in one case yields a new mechanistic hypothesis. Thus overall, this work not only yields insights into the computational mechanisms underlying the striking predictive capabilities of the retina, but also places the framework of deep networks as neuroscientific models on firmer theoretical founda- tions, by providing a new roadmap to go beyond comparing neural representations to extracting and understand computational mechanisms. 1 Introduction Deep convolutional neural networks (CNNs) have emerged as state of the art models of a variety of visual brain regions in sensory neuroscience, including the retina [1, 2], primary visual cortex (V1), [3, 4, 5, 6], area V4 [3], and inferotemporal cortex (IT) [3, 4]. Their success has so far been primarily evaluated by their ability to explain reasonably large fractions of variance in biological neural responses across diverse visual stimuli. However, fraction of variance explained is not of course the same thing as scientific explanation, as we may simply be replacing one inscrutable black box (the brain), with another (a potentially overparameterized deep network). Indeed, any successful scientific model of a biological circuit should succeed along three fundamental axes, each of which goes above and beyond the simple metric of mimicking the circuit's input- 33rd Conference on Neural Information Processing Systems (NeurIPS 2019), Vancouver, Canada. Figure 1: Deep learning models of the retina trained only on nat- ural scenes reproduce an array of retinal phenomena with artifi- cial stimuli (reproduced from ref. [2]). (A) Training procedure: We ana- lyzed a three-layer convolutional neural network (CNN) model of the retina which takes as input a spa- tiotemporal natural scene movie and outputs a nonnegative firing rate, corresponding to a retinal ganglion cell response. The first layer con- sists of eight spatiotemporal convo- lutional filters (i.e., cell types) with the size of (15×15×40), the second layer of eight convolutional filters (8×11×11), and the fully connected layer predicting the ganglion cells' response. As previously reported in [2], the deep learning model re- produces (B) an omitted stimulus re- sponse, (C) latency coding, (D) the motion reversal response, and (E) motion anticipation. output map. First, the intermediate computational mechanisms used by the hidden layers to generate responses should ideally match the intermediate computations in the brain. Second, we should be able to extract conceptual insight into how the neural circuit generates nontrivial responses to interesting stimuli (for example responses to stimuli that cannot be generated by a linear receptive field). And third, such insights should suggest new experimentally testable hypotheses that can drive the next generation of neuroscience experiments. However, it has been traditionally difficult to systematically extract computational mechanisms, and consequently conceptual insights, from deep CNN models due to their considerable complexity [7, 8]. Here we provide a method to do so based on the idea of model reduction, whose goal is to systematically extract a simple, reduced, minimal subnetwork that is most important in generating a complex CNN's response to any given stimulus. Such subnetworks then both summarize compu- tational mechanisms and yield conceptual insights. We build on ideas from interpretable machine learning, notably methods of input attribution that can decompose a neural response into a sum of contributions either from individual pixels [9] or hidden neurons [10]. To achieve considerable model reduction for responses to spatiotemporal stimuli, we augment and combine such input attribution methods with dimensionality reduction, which, for carefully designed artificial stimuli employed in neurophysiology experiments, often involves simple spatiotemporal averages over stimulus space. We demonstrate the power of our systematic model reduction procedure to attain mechanistic insights into deep CNNs by applying them to state of the art deep CNN models of the retina [1, 2]. The retina constitutes an ideal first application of our methods because the considerable knowledge (see e.g. [11]) about retinal mechanisms for transducing spatiotemporal light patterns into neural responses enables us to assess whether deep CNNs successfully learn the same computational structure. In particular, we obtain deep CNN models from [2] which were trained specifically to mimic the input-output transformation from natural movies to retinal ganglion cell outputs measured in the salamander retina. The model architecture involved a three-layer CNN model of the retina with ReLU nonlinearities (Fig. 1A). This network was previously shown [1, 2] to: (i) yield state of the art models of the retina's response to natural scenes that are almost as accurate as possible given intrinsic retinal stochasticity; (ii) possess internal subunits with similar response properties to those of retinal interneurons, such as bipolar and amacrine cell types; (iii) generalize from natural movies, to a wide range of eight different classes of artificially structured stimuli used over decades of neurophysiology experiments to probe 2 ΔtxyΔtxMotion reversalt [s]Schwartz et al. (2007)Rate [Hz]xLatency codingGollisch & Meister (2008)Figure 1Training input Natural scene movieTraining output Ganglion cells' responseExperimental dataOmitted stimulus responseModel outputSchwartz et al. (2007)sRate [Hz]t [s]Training procedure: natural scenesTesting procedure: structured stimulusMotion anticipationBerry et al. (1999)Position(A)(B)(C)(D)(E)N. Maheswaranathan et al. (2018) retinal response properties. This latter generalization capacity from natural movies to artificially structured stimuli (that were never present in the training data) is intriguing given the vastly different spatiotemporal statistics of the artificial stimuli versus natural stimuli, suggesting the artificial stimuli were indeed well chosen to engage the same retinal mechanisms engaged under natural vision [2]. Here, we focus on understanding the computational mechanisms underlying the deep CNN's ability to reproduce the neural responses to four classes of artificial stimuli (Fig. 1B-E), each of which, through painstaking experiments and theory, have revealed striking nonlinear retinal computations that advanced our scientific understanding of the retina. The first is the omitted stimulus response (OSR) [12, 13] (Fig. 1B), in which a periodic sequence of full field flashes entrains a retinal ganglion cell to respond periodically, but when a single flash is omitted, the ganglion cell produces an even larger response at the expected time of the response to the omitted flash. Moreover, the timing of this omitted stimulus response occurs at the expected time over a range of frequencies of the periodic flash train, suggesting the retina is somehow retaining a memory trace of the period of the train of flashes. The second is latency encoding [14], in which stronger stimuli yield earlier responses (Fig. 1C). The third is motion reversal [15], in which a bar suddenly reversing its motion near a ganglion cell receptive field generates a much larger response after the motion reversal (Fig. 1D). The fourth is motion anticipation [16], where the neural population responding to a moving bar is advanced in the direction of motion to compensate for propagation delays through the retina (Fig. 1E). These responses are striking because they imply the retina has implicitly built into it a predictive world model codifying simple principles like temporal periodicity, and the velocity based extrapolation of future position. The retina can then use these predictions to improve visual processing (e.g. in motion anticipation), or when these predictions are violated, the retina can generate a large response to signal that deviation (e.g. in the OSR and motion reversal). While experimentally motivated prior theoretical models have been employed to explain the OSR [17, 18], latency encoding [14, 19], motion reversal [20, 21], and motion anticipation [16], to date, no single model other than the deep CNN found in [2] has been able to simultaneously account for retinal ganglion cell responses to both natural scenes and all four of these classes of stimuli, as well as several other classes of artificial stimuli. However, it is difficult to explain the computational mechanisms underlying the deep CNN's ability to generate these responses simply by examining the complex network in Fig. 1A. For example, why does the deep CNN fire more when a stimulus is omitted, or when a bar reverses? How can it anticipate motion to compensate for propagation delays? And why do stronger responses cause earlier firing? These are foundational scientific questions about the retina whose answers require conceptual insights that are not afforded by the existence of a complex but highly predictive CNN alone. And even more importantly, if we could extract conceptual insights into the computational mechanisms underlying CNN responses, would these mechanisms match those used in the biological retina? Or is the deep CNN only accurate at the level of modelling the input-output map of the retina, while being fundamentally inaccurate at the level of underlying mechanisms? Adjudicating between these two possibilities is essential for validating whether the deep learning approach to modelling in sensory neuroscience can indeed succeed in elucidating biological neural mechanisms, which has traditionally been the gold-standard of circuit based understanding in systems neuroscience [11, 22, 23, 24]. In the following we will show how a combination of dimensionality reduction and hidden neuron or stimulus attribution can yield simplified subnetwork models of the deep CNNs response to stimuli, finding models that are consistent with prior mechanistic models with experimental support in the case of latency encoding, motion reversal, and motion anticipation. In addition, our analysis yields a new model that cures the inadequacies of previous models of the OSR. Thus our overall approach provides a new roadmap to extract mechanistic insights into deep CNN function, confirms in the case of the retina that deep CNNs do indeed learn computational mechanisms that are similar to those used in biological circuits, and yields a new experimentally testable hypothesis about retinal computation. Moreover, our results in the retina yield hope (to be tested in future combined theory and experiments) that more complex deep CNN models of higher visual cortical regions, may not only yield accurate black box models of input-output transformations, but may also yield veridical and testable hypotheses about intermediate computational mechanisms underlying these transformations, thereby potentially placing deep CNN models of sensory brain regions on firmer epistemological foundations. 3 2 From deep CNNs to neural mechanisms through model reduction To extract understandable reduced models from the millions of parameters comprising the deep CNN in Fig. 1A and [2], we first reduce dimensionality by exploiting spatial invariance present in the artificial stimuli carefully designed to specifically probe retinal physiology (Fig.1B-E), and then carve out important sub-circuits using modern attribution methods [9, 10]. We proceed in 3 steps: Step (1): Quantify the importance of a model unit with integrated gradients. The nonlinear input-output map of our deep CNN can be expressed as r(t) = F[s(t)], where r(t) ∈ R+ denotes the nonnegative firing rate of a ganglion cell at time bin t and s(t) ∈ R50×50×40 denotes the recent spatiotemporal history of the visual stimulus spanning two dimensions of space (x, y) (with 50 spatial bins in each dimension) as well as 40 preceding time bins parameterized by ∆t. Thus a single component of the vector s(t) is given by sxy∆t(t), which denotes the stimulus contrast at position (x, y) at time t − ∆t. We assume a zero contrast stimulus yields no response (i.e. F[0] = 0). We can decompose, or attribute the response r(t) to each preceding spacetime point by considering a straight path in spatiotemporal stimulus space from the zero stimulus to s(t) given by s(t; α) = αs(t) where the path parameter α ranges from 0 to 1 [9]. Using the line integral F[s(t; 1)] =(cid:82) 1 ∂α , we obtain ∂s(cid:12)(cid:12)s(t,α) · ∂s(t,α) 0 dα ∂F 50(cid:88)x=1 40(cid:88)∆t=1 50(cid:88)y=1 r(t) = F[s(t)] = sxy∆t(t)(cid:90) 1 0 dα ∂F ∂sxy∆t(t)(cid:12)(cid:12)(cid:12)(cid:12)αs(t) ≡ 50(cid:88)x=1 50(cid:88)y=1 40(cid:88)∆t=1 Axy∆t. (1) ∂s(t)(cid:12)(cid:12)s=0 · s(t) is only approximate. The coefficient vector ∂F This equation represents an exact decomposition of the response r(t) into attributions Axy∆t from each preceding spacetime stimulus pixel (x, y, ∆t). Intuitively, the magnitude of Axy∆t tells us how important each pixel is in generating the response, and the sign tells us whether or not turning on each pixel from 0 to sxy∆t(t) yields a net positive or negative contribution to r(t). When F is linear, this decomposition reduces to a Taylor expansion of F about s(t) = 0. However, in the nonlinear case, this decomposition has the advantage that it is exact, while the linear Taylor expansion r(t) ≈ ∂F of this Taylor expansion is often thought of as the linear spacetime receptive field (RF) of the model ganglion cell, a concept that dominates sensory neuroscience. Thus choosing to employ this attribution method enables us to go beyond the dominant but imperfect notion of an RF, in order to understand nonlinear neural responses to arbitrary spatiotemporal stimuli. In supplementary material, we discuss how this theoretical framework of attribution to input space can be temporally extended to answer different questions about how deep networks process spatiotemporal inputs. However, since our main focus here is model reduction, we consider instead attributing the ganglion cell response back to the first layer of hidden units, to quantify their importance. We denote by cxy(t) = W [1] z[1] cxy (cid:126) s(t) + bcxy the pre-nonlinearity activation of the layer 1 hidden units, where Wcxy and bcxy are the convolutional filters and biases of a unit in channel c (c = 1, . . . , 8) at convolutional position (x, y) (with x, y = 1, . . . , 36). Now computing the line integral F[s(t; 1)] = (cid:82) 1 0 dα ∂F ∂α over the same stimulus path employed in (1) yields ∂s(t)(cid:12)(cid:12)s=0 ∂z[1](cid:12)(cid:12)s(t,α) · ∂z[1] r(t) = (cid:88)x,y,c(cid:34)(cid:90) 1 dα 0 ∂F ∂z[1] cxy(cid:12)(cid:12)(cid:12)(cid:12)s(t,α)(cid:35) (W [1] cxy (cid:126) s) = (cid:88)x,y,c [Gcxy(s)] (W [1] cxy (cid:126) s) = (cid:88)x,y,c Acxy. (2) This represents an exact decomposition of the response r(t) into attributions Acxy from each subunit at the same time t (since all CNN filters beyond the first layer are purely spatial). This attribution further splits into a product of W [1] cxy (cid:126) s, reflecting the activity of that subunit originating from spatiotemporal filtering of the preceding stimulus history, and an effective stimulus dependent weight Gcxy(s) from each subunit to the ganglion cell, reflecting how variations in subunit activity z[1] cxy as the stimulus is turned on from 0 to s(t) yield a net impact on the response r(t). A positive (negative) effective weight indicates that increasing subunit activity along the stimulus path yields a net excitatory (inhibitory) effect on r(t). Step (2): Exploiting stimulus invariances to reduce dimensionality. The attribution of the re- sponse r(t) to first layer subunits in (2) still involves 8 × 36 × 36 = 10, 368 attributions. We can, however, leverage the spatial uniformity of artificial stimuli used in neurophysiology experiments to 4 reduce this dimensionality. For example, in the OSR and latency coding, stimuli are spatially uniform, implying W [1] cxy (cid:126) s is independent of spatial indices (x, y). Thus, we can reduce the number of attributions to the number of channels via c (cid:126) s ≡ W [1] r(t) = 8(cid:88)c=1(cid:18) 36(cid:88)x=1 36(cid:88)y=1 Gcxy(s)(cid:19) · (W [1] c (cid:126) s) ≡ 8(cid:88)c=1 Gc(s) · (W [1] c (cid:126) s) ≡ 8(cid:88)c=1 Ac. (3) For the moving bar in both motion reversal and motion anticipation, W [1] independent of the y index and we can reduce the dimensionality from 10,368 down to 288 by cxy (cid:126) s is r(t) = 8(cid:88)c=1 36(cid:88)x=1(cid:18) 36(cid:88)y=1 Gcxy(s)(cid:19) · (W [1] cx (cid:126) s) ≡ 8(cid:88)c=1 36(cid:88)x=1 cx (cid:126) s ≡ W [1] 36(cid:88)x=1 8(cid:88)c=1 Gcx(s) · (W [1] cx (cid:126) s) ≡ (4) Acx. More generally for other stimuli with no obvious spatial invariances, one could still attempt to reduce dimensionality by performing PCA or other dimensionality reduction methods on the space of hidden unit pre-activations or attributions over time. We leave this intriguing direction for future work. Step (3): Building reduced models from important subunits. Finally, we can construct minimal circuit models by first identifying "important" units defined as those with large magnitude attributions A. We then construct our reduced model as a one hidden layer neural network composed of only the important hidden units, with effective connectivity from each hidden unit to the ganglion cell determined by the effective weights G in (2), (3), or (4). 3 Results: the computational structure of retinal prediction We now apply the systematic model reduction steps described in the previous section to each of the retinal stimuli in Fig. 1B-E. We show that in each case the reduced model yields scientific hypotheses to explain the response, often consistent with prior experimental and theoretical work, thereby validating deep CNNs as a method for veridical scientific hypothesis generation in this setting. Moreover, our approach yields integrative conceptual insights into how these diverse computations can all be simultaneously produced by the same set of hidden units. 3.1 Omitted stimulus response As shown in Fig. 1B, periodic stimulus flashes trigger delayed periodic retinal responses. However, when this periodicity is violated by omitting a flash, the ganglion cell signals the violation with a large burst of firing [25, 26]. This OSR phenomenon is observed across several species including salamander [12, 13]. Interestingly, for periodic flashes in the range of 6-12Hz, the latency between the last flash before the omitted one, and the burst peak in the response, is proportional to the period of the train of flashes [12, 13], indicating the retina retains a short memory trace of this period. Moreover, pharmacological experiments suggest ON bipolar cells are required to produce the OSR [13, 17], which have been shown to correspond to the first layer hidden units in the deep CNN [1, 2]. These phenomena raise two fundamental questions: what computational mechanism causes the large amplitude burst, and how is the timing of the peak sensitive to the period of the flashes? There are two theoretical models in the literature that aim to answer these questions. One proposes that the bipolar cell activity responds to each individual flash with an oscillatory response whose period adapts to the period of the flash train [18]. However, recent direct recordings of bipolar cells suggest that such period adaptation is not present [27]. The other model claims that having dual pathways of ON and OFF bipolar cells are enough to reproduce most of the aspects of the phenomena observed in experiments [17]. However, the model only reproduces the shift of the onset of the burst, and not a shift in the peak of the burst, which has the critical predictive latency [18]. Direct model reduction (Fig. 2) of the deep CNN in Fig. 1A using the methods of section 2 yields a more sophisticated model than any prior model, comprised of three important pathways that combine one OFF temporal filter with two ON temporal filters. Unlike prior models, the reduced model exhibits a shift in the peak of the OSR burst as a function of the frequency of input flashes. 5 Figure 2: Omitted stimulus response. (A-1,2) Schematics of the model reduction procedure by only leaving three (1 OFF, 2 ON) highly contributing units. (B) Attribution for each of the cell types Ac over time. (C) Effective stimulus dependent weight for each of the cell types Gc over time. (D) The combination of the two pathways of filter 2 and 6 reproduces the period dependent latency. (E) Two ON bipolar cells are necessary to capture the predictive latency. Cell 2 with earlier peak is only active in a high-frequency regime, while the cell 6 with later peak is active independent of the frequency. Fig. 2A presents a schematics of the model reduction steps described in (3). We first attribute a ganglion cell's response to 8 individual channels and then average across both spatial dimensions (Fig. 2A-1) as in steps (1) and (2). Then we build a reduced model from the identified important subunits that capture essential features of the omitted stimulus response phenomenon (Fig. 2A-2). In Fig. 2B, we present the time dependence of the attribution Ac(s(t)) in (3) for the eight channels, or cell-types. Red (blue) traces reflect positive (negative) attributions. Channel temporal filters are to the left of each attribution row and the total output response r(t) is on the top row. The stimulus consists of three flashes, yielding three small responses, and then one large response after the end of the three flashes (grey line). Quantitatively, we identify that cell-type 3 dominantly explains the small responses to preceding flashes, while cell types 2 and 6 are necessary to explain the large burst after the flash train ends. The final set of units included in the reduced model should be the minimal set required to capture the defining features of the phenomena of interest. In the case of omitted stimulus response, the defining feature is the existence of the large amplitude burst whose peak location is sensitive to the period of the applied flashes. Once we identify the set of essential temporal filters, we then proceed to determine the sign and magnitude of contribution (excitatory or inhibitory) of the cell types. In Fig. 2C, we present the time-dependent effective weights from Gc(s(t)) in (3) for the eight cell types, or channels. Red (blue) reflects positive (negative) weights. Given the product of the temporal filters and the weights, cell-types 2 and 6 are effectively ON cells, which cause positive ganglion cell responses to contrast increments, while cell-type 3 is an OFF cell, which is a cell type that causes positive responses to contrast decrements. Following the prescribed procedures, carving out the 3 important cell-types and effective weights yields a novel, mechanistic three pathway model of the OSR, with 1 OFF and 2 ON pathways. Unlike prior models, the reduced model exhibits a shift in the peak of the OSR burst as a function of the frequency of input flashes (with dark to light blue indicating high to low frequency variation in the flash train) as in Fig. 2D. Furthermore, the reduced model is consistent across the frequency range that produces the phenomena. Finally, model reduction yields conceptual insights into how cell-types 2 and 6 enable the timing of the burst peak to remember the period of the flash train (Fig. 2E). The top row depicts the decomposition of the overall burst response r(t) (grey) into time dependent attributions A2 (red) and A6 (blue), obeying the relation r(t) ≈ A2 + A6. Cell-type 2, which has an earlier peak in its temporal filter, preferentially causes ganglion cell responses in high-frequency flash trains (left) compared to low frequency trains (right), while cell-type 6 is equally important in both. The middle row shows the temporal filter Wc=2(∆t), which has an earlier peak with a long tail, enabling it to integrate across not only the last flash, but also preceding flashes (yellow bars). Time increases into the past from left to right. Thus, the activation of this cell type 2 decreases as the flash train frequency decreases, explaining the decrease in attribution in the top row. The bottom row shows that the temporal filter Wc=6(∆t) of cell type 6, in contrast, has a later peak with a rapidly decaying tail. Thus the temporal convolution Wc=6(∆t) (cid:126) s(∆t) of this filter with the flash train is sensitive only to the last flash, and 6 (A-2)Ac=6(t)timeintensityFigure 2: Omitted Stimulus Response(A-1)(B)(C)~(D)s(t)(E)12345678Ac=Gc·(W[1]c~s)r(t)=8Xc=1AcGcGc(s)W[1]c32612345678high freqlow freq26Ac=2(t)r(t)⇡+tttW[1]c=2~sW[1]c=6~shigh freqlow freqPositive Negative is therefore independent of flash train frequency. The late peak and rapid tail explain why it supports the response at late times independent of frequency in the top row. Thus, our systematic model reduction approach yields a new model of the OSR that cures important inadequacies of prior models. Moreover, it yields a new, experimentally testable scientific hypothesis that the OSR is an emergent property of three bipolar cell pathways with specific and diverse temporal filtering properties. 3.2 Latency coding Rapid changes in contrast (which often occur for example right after saccades) elicit a burst of firing in retinal ganglion cells with a latency that is shorter for larger contrast changes [14] (Fig. 1C). Moreover, pharmacological studies demonstrate that both ON and OFF bipolar cells (corresponding to first layer hidden neurons in the deep CNN [1, 2]) are necessary to produce this phenomenon [19]. Model reduction via (3) in section 2 reveals that a single pair of slow ON and fast OFF pathways can explain the shift in the latency Fig. 3. First, under a contrast decrement, there is a strong, fast excitatory contribution from the OFF pathway. Second, as the magnitude of the contrast decre- ment increases, delayed inhibition from the slow ON pathway becomes stronger. This negative delayed contribution truncates excitation from the OFF pathway at late times, thereby causing a shift in the location of the total peak response to earlier times (Fig. 3). The dual pathway mech- anism formed by slow ON and fast OFF bipolar cells is consistent with all existing experimental facts. Moreover, it has been previously proposed as a theory of latency coding [14, 19]. Thus this example illustrates the power of a general nat- ural scene based deep CNN training approach, followed by model reduction, to automatically generate veridical scientific hypotheses that were previously discovered only through specialized experiments and analyses requiring significant effort [14, 19]. Figure 3: Latency coding (A) The decomposition of the overall response r(t) (grey) into dominant attributions A3(t) (blue) from an OFF pathway, and A2(t) (red) from an ON pathway, obeying the relation r(t) ≈ A3 + A2. Under a contrast decre- ment, the OFF pathway activated first, followed by delayed inhibitory input from the ON pathway. (B) As the amount of contrast decrement increases (yel- low bars), delayed inhibition from the ON pathway (red) strengthens, which cuts off the total response in r(t) at late times more strongly, thereby shifting the location of the peak of r(t) to earlier times. 3.3 Motion reversal As shown in Fig. 1D and [15], when a moving bar suddenly reverses its direction of motion, ganglion cells near the reversal location exhibit a sharp burst of firing. While a ganglion cell classically responds as the bar moves through its receptive field (RF) center from left to right before the motion reversal, the sharp burst after the motion reversal does not necessarily coincide with the spatial re-entry of the bar into the center of the RF as it moves back from right to left. Instead, the motion reversal burst response occurs at a fixed temporal latency relative to the time of motion reversal, for a variety of reversal locations within 110 µm of the RF center. These observations raise two fundamental questions: why does the burst even occur and why does it occur at a fixed latency? The classical linear-nonlinear model cannot reproduce the reversal response; it only correctly re- produces the initial peak associated with the initial entry of a bar into the RF center [15]. Thus a nonlinear mechanism is required. Model reduction of the deep CNN obtained via (4) reveals that two input channels arrayed across 1D x space can explain this response through a specific nonlinear mechanism (Fig. 4). Moreover, the second important channel revealed by model reduction yields a cross cell-type inhibition that explains the fixed latency (Fig. 4D). Intriguingly, this reduced model is 7 Figure 3: Latency Coding23inputoutputLow intensity(A)(B)High intensityA2A3r(t)⇡+ Figure 4: Motion reversal of a moving bar. (A) Schematics of (x, t) spatiotemporal model reduction obtained via (4). By averaging over y we obtain 8 cell types at 36 different x positions yielding 288 units. Attribution values reveal that only cell types 2 and 3 play a dominant role in motion reversal. (B) The properties of cell-type 3 explains the existence of the burst. On the left are time-series of pre-nonlinearity activations W [1] c=3,x (cid:126) s of hidden units whose RF center is at spatial position x. Time t = 0 indicates the time of motion reversal. The boxed region indicates the spatial and temporal extent of the retinal burst in response to motion reversal. The offset of the box from time t = 0 indicates the fixed latency. A fixed linear combination with constant coefficents of this activation cannot explain the existence of the burst due to cancellations along the vertical x-axis in the boxed region. However, due to downstream nonlinearities, the effective weight coefficients Gc=3,x from subunits to ganglion cell responses rapidly flip in sign (middle), and generating a burst of motion reversal response (right). (C) Schematics of the reduced model keeping only important subunits. (D) Attribution contributions from the two dominant cell types A2 (in pink) and A3 (in blue), where Ac =(cid:80)36 x=1 Acx. With only cell-type 3, the further the reversal location is from a ganglion cell's RF center, the longer we would expect it to take to generate a reversal response. However, the inhibition coming from cell type 2 increases the further away the reversal occurs, truncating the late response and thus fixing the latency of the motion reversal response. qualitatively consistent with a recently proposed and experimentally motivated model [20] that points out the crucial role of dual pathways of ON and OFF bipolar cells. 3.4 Motion anticipation As shown in Fig. 1E and [16] the retina already starts to compensate for propagation delays by advancing the retinal image of a moving bar along the direction of motion, so that the retinal image does not lag behind the instantaneous location as one might naively expect. Model reduction of our deep CNN reveals a mechanism for this predictive tracking. First, since ganglion cell RFs have some spatial extent, a moving bar naturally triggers some ganglion cells before entering their RF center, yielding a leading edge of a retinal wave. What is then required for motion anticipation is some addi- tional motion direction sensitive inhibition that cuts off the lagging edge of the wave so its peak activity shifts towards the leading edge. Indeed, model reduction reveals a computational mech- anism in which one cell type feeds an excitatory signal to a ganglion cell while the other provides direction sensitive inhibition that truncates the lagging edge. This model is qualitatively consistent with prior theoretical models that employ such direction selective inhibition to anticipate motion [16]. Figure 5: Motion anticipation of a moving bar. Contributions from the two dominant cell types. A2 in pink, A3 in blue, r(t) ≈ A2 + A3 in grey, where Ac =(cid:80)36 x=1 Acx. Depending on the direc- tion of motion of a bar, activity that lags behind the leading edge gets asymmetrically truncated by the inhibition from the cell type 2 (pink). (A) The bar is moving to the right and the inhibition (pink) is slightly stronger on the left side. (B) the bar is moving to the left and the inhibition (pink) is stronger on the right side. 8 Figure 4: Motion Reversal(A)(B)(D)(C)txW[1]cx~sGcxr(t)=8Xc=136Xx=1AcxW[1]c=3,x~sGc=3,xAc=3,x=Gc=3,x·(W[1]c=3,x~s)Ac=3,x=Gc=3,x·(W[1]c=3,x~s)xA2A3r(t)⇡+32Figure 5: Motion Anticipation(A)(B)A2A3r(t)⇡+ 4 Discussion Figure 6: A unified framework to reveal computational structure in the brain. We outlined an auto- mated procedure to go from large- scale neural recordings to mechanis- tic insights and scientific hypotheses through deep learning and model re- duction. We validate our approach on the retina, demonstrating how only three cell-types with different ON/OFF and fast/slow spatiotempo- ral filtering properties can nonlin- early interact to simultaneously gen- erate diverse retinal responses. In summary, in the case of the retina, we have shown that complex CNN models obtained via machine learning can not only mimic sensory responses to rich natural scene stimuli, but also can serve as a powerful and automatic mechanism for generating valid scientific hypotheses about computational mechanisms in the brain, when combined with our proposed model reduction methods (Fig. 6). Applying this approach to the retina yields conceptual insights into how a single model consisting of multiple nonlinear pathways with diverse spatiotemporal filtering properties can explain decades of painstaking physiological studies of the retina. This suggests in some sense an inverse roadmap for experimental design in sensory neuroscience. Rather than carefully designing special artificial stimuli to probe specific sensory neural responses, and generating individual models tailored to each stimulus, one could instead fit a complex neural network model to neural responses to a rich set of ethologically relevant natural stimuli, and then apply model reduction methods to understand how different parts of a single model can simultaneously account for responses to artificial stimuli across many experiments. The interpretable mechanisms extracted from model reduction then constitute specific hypotheses that can be tested in future experiments. Moreover, the complex model itself can be used to design new stimuli, for example by searching for stimuli that yield divergent responses in the complex model, versus a simpler model of the same sensory region. Such stimulus searches could potentially elucidate functional reasons for the existence of model complexity. In future studies, it will be interesting to conduct a systematic exploration of universality and individuality [28] in the outcome of model reduction procedures applied to deep learning models which recapitulate desired phenomena, but are obtained from different initializations, architectures, and experimental recordings. An intriguing hypothesis is that the reduced models required to explain specific neurobiological phenomena arise as universal computational invariants across the ensemble of deep learning models parameterized by these various design choices, while many other aspects of such deep learning models may individually vary across these choices, reflecting mere accidents of history in initialization, architecture and training. It would also be extremely interesting to stack this model reduction procedure to obtain multilayer reduced models that extract computational mechanisms and conceptual insights into deeper CNN models of higher cortical regions. The validation of such extracted computational mechanisms would require further experimental probes of higher responses with carefully chosen stimuli, perhaps even stimuli chosen to maximize responses in the deep CNN model itself [29, 30]. Overall the success of this combined deep learning and model reduction approach to scientific inquiry in the retina, which was itself not at all a priori obvious before this work, sets a foundation for future studies to explore this combined approach deeper in the brain. Acknowledgments We thank Daniel Fisher for insightful discussions and support. We thank the Masason foundation (HT), grants from the NEI (R01EY022933, R01EY025087, P30-EY026877) (SAB), and the Simons, James S. McDonnell foundations, and NSF Career 1845166 (SG) for funding. 9 1. High-throughput neural recordings2. Train deep-learning model and perform neurophysiology experiments in silico3. Identify important sub-circuits and derive an array of interpretable models236Figure 6 Final References [1] Lane McIntosh, Niru Maheswaranathan, Aran Nayebi, Surya Ganguli, and Stephen Baccus. Deep learning models of the retinal response to natural scenes. Advances in neural information processing systems, pages 1369 -- 1377, 2016. [2] Niru Maheswaranathan, Lane T McIntosh, David B Kastner, Josh Melander, Luke Brezovec, Aran Nayebi, Julia Wang, Surya Ganguli, and Stephen A Baccus. Deep learning models reveal internal structure and diverse computations in the retina under natural scenes. bioRxiv, page 340943, 2018. [3] Daniel L. K. Yamins, Ha Hong, Charles F. Cadieu, Ethan A. Solomon, Darren Seibert, and James J. DiCarlo. Performance-optimized hierarchical models predict neural responses in higher visual cortex. Proceedings of the National Academy of Sciences, 111(23):8619 -- 8624, 2014. doi: 10.1073/pnas.1403112111. [4] Seyed-Mahdi Khaligh-Razavi and Nikolaus Kriegeskorte. Deep supervised, but not unsupervised, models may explain it cortical representation. PLoS computational biology, 10(11):e1003915, 2014. [5] Umut Güçlü and Marcel AJ van Gerven. Deep neural networks reveal a gradient in the complexity of neural representations across the ventral stream. The Journal of Neuroscience, 35(27):10005 -- 10014, 2015. [6] Santiago A Cadena, George H Denfield, Edgar Y Walker, Leon A Gatys, Andreas S Tolias, Matthias Bethge, and Alexander S Ecker. Deep convolutional models improve predictions of macaque v1 responses to natural images. bioRxiv, page 201764, 2017. [7] David GT Barrett, Ari S Morcos, and Jakob H Macke. Analyzing biological and artificial neural networks: challenges with opportunities for synergy? Current opinion in neurobiology, 55:55 -- 64, 2019. [8] Joshua I Glaser, Ari S Benjamin, Roozbeh Farhoodi, and Konrad P Kording. The roles of supervised machine learning in systems neuroscience. Progress in neurobiology, 2019. [9] Mukund Sundararajan, Ankur Taly, and Qiqi Yan. Axiomatic attribution for deep networks. In Doina Precup and Yee Whye Teh, editors, Proceedings of the 34th International Conference on Machine Learning, volume 70 of Proceedings of Machine Learning Research, pages 3319 -- 3328, International Convention Centre, Sydney, Australia, 06 -- 11 Aug 2017. PMLR. URL http://proceedings.mlr.press/v70/ sundararajan17a.html. [10] Kedar Dhamdhere, Mukund Sundararajan, and Qiqi Yan. How important is a neuron. In International Con- ference on Learning Representations, 2019. URL https://openreview.net/forum?id=SylKoo0cKm. [11] Tim Gollisch and Markus Meister. Eye smarter than scientists believed: neural computations in circuits of the retina. Neuron, 65(2):150 -- 164, 2010. [12] Greg Schwartz, Rob Harris, David Shrom, and Michael J Berry II. Detection and prediction of periodic patterns by the retina. Nature neuroscience, 10(5):552, 2007. [13] Greg Schwartz and Michael J Berry 2nd. Sophisticated temporal pattern recognition in retinal ganglion cells. Journal of neurophysiology, 99(4):1787 -- 1798, 2008. [14] Tim Gollisch and Markus Meister. Rapid neural coding in the retina with relative spike latencies. science, 319(5866):1108 -- 1111, 2008. [15] Greg Schwartz, Sam Taylor, Clark Fisher, Rob Harris, and Michael J Berry II. Synchronized firing among retinal ganglion cells signals motion reversal. Neuron, 55(6):958 -- 969, 2007. [16] Michael J Berry II, Iman H Brivanlou, Thomas A Jordan, and Markus Meister. Anticipation of moving stimuli by the retina. Nature, 398(6725):334, 1999. [17] Birgit Werner, Paul B Cook, and Christopher L Passaglia. Complex temporal response patterns with a simple retinal circuit. Journal of neurophysiology, 100(2):1087 -- 1097, 2008. [18] Juan Gao, Greg Schwartz, Michael J Berry, and Philip Holmes. An oscillatory circuit underlying the detection of disruptions in temporally-periodic patterns. Network: Computation in Neural Systems, 20(2): 106 -- 135, 2009. [19] Tim Gollisch and Markus Meister. Modeling convergent on and off pathways in the early visual system. Biological cybernetics, 99(4-5):263 -- 278, 2008. [20] Eric Y Chen, Janice Chou, Jeongsook Park, Greg Schwartz, and Michael J Berry. The neural circuit mechanisms underlying the retinal response to motion reversal. Journal of Neuroscience, 34(47):15557 -- 15575, 2014. 10 [21] Eric Y Chen, Olivier Marre, Clark Fisher, Greg Schwartz, Joshua Levy, Rava Azeredo da Silveira, and Michael J Berry. Alert response to motion onset in the retina. Journal of Neuroscience, 33(1):120 -- 132, 2013. [22] David I Vaney, Benjamin Sivyer, and W Rowland Taylor. Direction selectivity in the retina: symmetry and asymmetry in structure and function. Nature Reviews Neuroscience, 13(3):194, 2012. [23] Masakazu Konishi. Coding of auditory space. Annual review of neuroscience, 26(1):31 -- 55, 2003. [24] Eve Marder and Dirk Bucher. Understanding circuit dynamics using the stomatogastric nervous system of lobsters and crabs. Annu. Rev. Physiol., 69:291 -- 316, 2007. [25] Theodore H Bullock, Michael H Hofmann, Frederick K Nahm, John G New, and James C Prechtl. Event- related potentials in the retina and optic tectum of fish. Journal of Neurophysiology, 64(3):903 -- 914, 1990. [26] Theodore H Bullock, Sacit Karamürsel, Jerzy Z Achimowicz, Michael C McClune, and Canan Ba¸sar- Eroglu. Dynamic properties of human visual evoked and omitted stimulus potentials. Electroencephalog- raphy and clinical neurophysiology, 91(1):42 -- 53, 1994. [27] Nikhil Rajiv Deshmukh. Complex computation in the retina. PhD thesis, Princeton University, 2015. [28] Niru Maheswaranathan, Alex H Williams, Matthew D Golub, Surya Ganguli, and David Sussillo. Univer- sality and individuality in neural dynamics across large populations of recurrent networks. Advances in neural information processing systems, 2019. [29] Pouya Bashivan, Kohitij Kar, and James J DiCarlo. Neural population control via deep image synthesis. Science, 364(6439):eaav9436, 2019. [30] Edgar Y Walker, Fabian H Sinz, Emmanouil Froudarakis, Paul G Fahey, Taliah Muhammad, Alexander S Ecker, Erick Cobos, Jacob Reimer, Xaq Pitkow, and Andreas S Tolias. Inception in visual cortex: in vivo-silico loops reveal most exciting images. bioRxiv, page 506956, 2018. [31] Robert J Aumann and Lloyd S Shapley. Values of non-atomic games. Princeton University Press, 1974. 11 Supplemental Materials 5 Path-integrated gradients for nonlinear dynamical systems A specific firing burst of a ganglion cell r(t) (e.g. omitted stimulus response, motion reversal etc.) is a result of spatiotemporal visual stimuli s(t, x, y) processed through a non-linearly entangled web of neurons, r(t) = F[s(t, x, y)]. Here we aim to identify and isolate the minimal neural circuit at play, in order to obtain theoretical and biophysical insights responsible for an array of retinal phenomena. Towards this goal, we need to attribute the contributions to each of the input pixels or interneurons. For a linear system, the product of the response function and an input stimulus χ(∆t, x, y)s(t − ∆t, x, y) quantifies how much an input pixel changes the output. However, it is not obvious how to introduce such measure local in (t, x, y) space for non-linear systems, thus raises questions: how can we, in non-linear systems, quantify the contributions of each of the (1) spatiotemporal input pixels, (2) internal neurons, and (3) different cell types (filters)? Here we harness attribution methods based on path-integrated gradients, first introduced in game theory [31], and then recently in machine learning [9]. We then extend the theoretical framework to a movie input with an additional dimension of time. Retina as a non-linear functional Mathematically, a mapping from a space of spatiotemporal stimulus s(t, x, y) to a scalar response of a ganglion cell r(t) at time t can be represented by a functional, (5) Here we assumed (i) causality and (ii) finite temporal memory of a ganglion cell with a cutoff at ∆tm. Thus, the input s(∆t, x, y; t) is a subset of the entire spatiotemporal stimulus s(t(cid:48), x, y) in a finite region t ∈ [t − ∆tm, t] and parametrized by t. Expansion of a general functional v.s. linear response theory The Taylor expansion of the above functional around s(∆t, x, y; t) ∼ 0 is, r(t) = F(cid:2)s(∆t, x, y; t)(cid:3). r(t) = F(cid:2)s(∆t, x, y; t)(cid:3) (cid:90) = F [0] + d(∆t1)dx1dy1 δs(∆t1, x1, y1; t) In comparison, the linear response theory takes the form of, r(t) = F(cid:2)s(∆t, x, y; t)(cid:3) = F[0] + (cid:12)(cid:12)(cid:12)(cid:12)s=0 δF(cid:2)s(t)(cid:3) (cid:90) s(∆t1, x1, y1; t) + O(s2) d(∆t)dxdyχ(∆t, x, y)s(∆t, x, y; t), (6) (7) where χ is the linear response function (e.g. receptive field around zero stimuli). Two main approximations are made in the linear response theory; (i) no stimulus s dependence of δF δs , (ii) trun- cation of the Taylor expansion at the first order. These two assumptions are valid only when s(∆t, x, y; t) (cid:28) 1 holds everywhere. However, this condition hardly holds in visual systems in action, and the above conventional formulation is due to theoretical and experimental limitations rather than a well-justified approximation as in Figure 7A for schematics. Figure 7: How important is an input pixel or a cell type (filter)? (A) [Linear theory: red] Conventional linear theory, such as classical receptive field, extrapolates linear trend to all over the input stimulus space. Deep learning model trained on natural scenes is able to compute local gradient, "instantaneous receptive field", given an input of structured stimuli with high efficiency. However, the instantaneous receptive field only explains the local property of the non-linear function that does not necessalily reflect the global structure. [IntegratedGradients in 1D: blue] By integrating the local gradients captured by a deep learning model over input axis, we can obtain a more acculate description of the non-linear function. (B) The path-integrated gradients becomes dependents on the path in higher dimensions (D > 1), but straight path is shown to satisfy desired axioms [9]. Schematic illustrates the case of two-dimensions. 12 @F(↵x0,↵y0)@x@F(↵x0,↵y0)@y(x0,y0)F(x,y)(A)xx0rr0=dF(0)dx·x0(B)Instantaneous receptive field1D2Dr0=F(x0)=x0Z10d↵dF(↵x0)dxFigure 2: IntegratedGradientss0s1G While both of the biological retina and the deep learning model gives r(t) = F(cid:2)s(∆t, x, y; t)(cid:3) upon a presentation of a stimulus, the stimulus-dependent instantaneous receptive field (8) is easily computable in the deep learning model. Below, we harness the quantity δF/δss to investigate where a ganglion cell is looking at during a burst of neural firings. Review: Path integrated gradients for an image input Here we first review a case of an image input r = F[s(x, y)], where s(x, y) represents a two dimensional image. For a general path γ(α) parameterized by α, r(α = 1) − r(α = 0) = F[s(x, y, α = 1)] − F [s(x, y, α = 0)] (cid:90) (cid:20) δF(x, y; α) (cid:90) 1 0 = dxdy dα δs(x, y; α) ∂α (cid:21) (cid:90) ∂s(x, y; α) ≡ dxdyAγ(α)(x, y). (9) Sundararajan et al. [9] proved that a straight path s(x, y; α) = αs(x, y) satisfies an array of desirable axioms: Sensitivity, Insensitivity, Linearity, Preservation, Implementation invariance, Completeness, and Symmetry. In that case, the above further simplifies to, AIG(x, y) = s(x, y) F(x, y; α) δs(x, y; α) dα . (10) See Figure 7 for schematics. Space-time attribution by path integrated gradients for a movie input Here we try to extend the theoretical framework of the above IntegratedGradients to the case of a movie input. In addition to the spatial components (x, y), a movie input has an additional dimension of time t. In this case, the attribution score A(∆t, x, y; t) can be computed by taking a path integral parametrized by α, (cid:12)(cid:12)(cid:12)(cid:12)s δF δs (cid:90) (cid:90) (cid:90) d(∆t)dxdy dα (cid:20)(cid:90) r(t) = F(cid:2)s(∆t, x, y; t)(cid:3) − F(cid:2)s(∆t, x, y; α = 0)(cid:3) δF(cid:2)s(∆t, x, y; α)(cid:3) (cid:12)(cid:12)(cid:12)(cid:12)s(α) (cid:12)(cid:12)(cid:12)(cid:12)s(α) δF(cid:2)s(∆t, x, y; α)(cid:3) d(∆t)dxdyA(∆t, x, y; t). δs(∆t, x, y; α) δs(∆t, x, y; α) (cid:20)(cid:90) = ≡ dα (cid:21) (cid:21) . s(∆t, x, y; α) ∂α s(∆t, x, y; α) ∂α Thus, attribution score can be defined as A(∆t, xx, y; t) ≡ (11) (12) For a linear system, where δF δs is independent of s, we can simply replace it by a linear response function δs s=0 = χ, and the attribution score reduces to a product of the linear response function and the difference δF between the current stimulus and the baseline, A(∆t, x, y; t) = χ(∆t, x, y; t)(cid:0)s(∆t, x, y; t) − s(∆t, x, y; α = 0)(cid:1). (13) δs s depends on s and the attribution score A(∆t, x, y) depends on the However, for a non-linear system, δF space of function that we integrate over. Namely, each "path" corresponds to a different attribution method and provides different explanations to different questions. In particular, for a movie input, the history of stimuli forms a path naturally parametrized by time "t". IntegratedGradients: Effective linear response function for non-linear systems Here we focus on the straight path, s(∆t, x, y; α) = αs(∆t, x, y; t). (14) In this case, the integral of the gradients act as an effective linear response function for a nonlinear system as below, AIG(∆t, x, y; t) ≡ dα (cid:20)(cid:90) δF(cid:2)αs(∆t, x, y; t)(cid:3) δ(cid:0)αs(∆t, x, y; t)(cid:1) (cid:21) = χ(∆t, x, y; t)s(∆t, x, y; t) = χ[s](t, t (cid:48) (cid:48) , x, y)s(t , x, y). s(∆t, x, y; t) (15) By computing the attribution A(∆t, x, y; t) over the full temporal region, we can obtain the full attribution A(t, t(cid:48), x, y). Note that the function now depends both on t and t(cid:48) due to nonlinearity. Furthermore, we can reformulate the above equation into two forms to answer specific questions raised below. 13 input stimuli(cid:82) ∞ Figure 8: Straight path integration from the null stimulus We computed IntegratedGradients for a periodic flashes with omitted stimulus by taking a straight path integration of gradients from the null stimulus s0 = 0 to an actual stimulus s1. (A) (Left) Attribution map on response time t v.s input time t(cid:48) space. Causality limits the attribution in the region of t(cid:48) < t. (Right) Attribution on t(cid:48) dtA(t, t(cid:48)) quantifies the total amount of output created by each of the pixels of an input movie. (B) (Left) Attribution map on response time t v.s. integration time ∆t space. (Right) Attribution on integration time ∆t integrated over a single periodic burst triggered by flashes (blue) and omitted stimulus response (red). The peak of attribution for the flash response (blue) is placed earlier than the peak of attribution for the omitted stimulus response (orange). This is consistent with our finding that the responses for periodic flashes are caused by a fast OFF bipolar cell, while the omitted stimulus response is triggered by slow ON bipolar cells. (See main text.) the attribution in the region of t(cid:48) < t. (Right) Attribution on input stimuli(cid:82) ∞ Figure 9: Straight path integration between two stimuli We computed IntegratedGradients to discriminate between two stimuli continuing periodic flashes s0, and periodic flashes with omitted stimulus s1. (A) (Left) Attribution map on response time t v.s input time t(cid:48) space. Causality limits t(cid:48) dtA(t, t(cid:48)) quantifies the total amount of output created by each of the pixels of ano input movie. Strikingly, the attribution concentrates at the location of the first omitted stimulus among the four omitted stimulus. (B) (Left) Attribution map on response time t v.s. integration time ∆t space. (Right) Attribution on integration time ∆t integrated over a single periodic burst triggered by flashes (blue) and omitted stimulus response (orange). Q1. Which pixel in the spatiotemporal input stimulus "(t, x, y)" is responsible for a ganglion cell's re- sponse? (cid:90) t1 (cid:90) (cid:18)(cid:90) t1 dtr(t) = (cid:48) dt dxdy dtA(t, t (cid:48) , x, y) t0 t0 (cid:19) See Figure 8 for an example with OSR. Q2. Which part of the spatiotemporal kernel is "(∆t, x, y)" responsible for a ganglion cell's response? (16) (cid:19) (cid:90) t1 (cid:90) (cid:90) t1 (cid:90) t1 dtr(t) = dxdy dt t0 t0 t0−∆tm See Figure 9 for an example with OSR. , x, y)δ(cid:0)(t−t (cid:48) )−∆t(cid:1) = (cid:90) (cid:48)A(t, t dt (cid:48) (cid:18)(cid:90) t1 d∆tdxdy dtA(t, t−∆t, x, y) t0 (17) 6 Extension to deeper CNNs without spatial invariances in stimuli In the main text, we demonstrated how we can reduce CNNs with three layers to multi-pathways linear non-linear models, specifically to validate deep CNNs in the retina where we had neurobiological ground truth. Here, we discuss one possible way to extend our method to deeper CNNs of depth D processing natural movies through a dynamic programming (DP) approach that works backwards from layer D to layer 1. First, note a natural movie of limited duration without spatial invariances is still well approximated by a low dimensional trajectory in both pixel space and every hidden layer. Let K be the max dimensionality for spatial input patterns for any channel 14 SI Figure: Path-integrated gradients for spatiotemporal inputOSRFlashesOSRFlashess0s1Straight path integration from the null stimuliStraight path integration between stimulus, and s0s1(A)(B)(A)(B)s0s1SI Figure: Path-integrated gradients for spatiotemporal inputOSRFlashesOSRFlashess0s1Straight path integration from the null stimuliStraight path integration between stimulus, and s0s1(A)(B)(A)(B)s0s1 in any layer. Then the basic idea is to attribute the response in layer D to the K dimensional space of inputs to each channel in layer D − 1 using integrated gradients. We first find the important channels in layer D − 1 using methods presented in the main text. Then we recursively iterate via the same method to layer D − 2 and so on down to the pixel layer. Because of the DP-like nature of our algorithm, the computational complexity (after dimensionality reduction to K) is O(DKC) where C is the max number of channels in a layer, and not exponential in D. The end result is a set of important channels in each layer, along with, for each important channel ≤ K linear combinations of neurons that matter for generating the response in layer D. 15
1801.09900
1
1801
2018-01-30T09:01:00
Human Echolocation in Static Situations: Auditory Models of Detection Thresholds for Distance, Pitch, Loudness and Timbre
[ "q-bio.NC" ]
We investigated, by using auditory models, how three perceptual parameters, loudness, pitch and sharpness, determine human echolocation. We used acoustic recordings from two previous studies, both from stationary situations, and their resulting perceptual data as input to our analysis. An initial analysis was on the room acoustics of the recordings. The parameters of interest were sound pressure level, autocorrelation and spectral centroid. The auditory models were used to analyze echolocation resulting from the perceptual variables, i.e. loudness, pitch and sharpness. Relevant auditory models were chosen to simulate each variable. Based on these results, we calculated psychophysical thresholds for detecting a reflecting object with constant physical size. A non-parametric method was used to determine thresholds for distance, loudness, pitch and sharpness. Difference thresholds were calculated for the psychophysical variables, since a 2-Alternative-Forced-Choice Paradigm had originally been used. We found that (1) blind persons could detect objects at lower loudness values, lower pitch strength, different sharpness values and at further distances than sighted persons, (2) detection thresholds based on repetition pitch, loudness and sharpness varied and depended on room acoustics and type of sound stimuli, (3) repetition pitch was useful for detection at shorter distances and was determined from the peaks in the temporal profile of the autocorrelation function, (4) loudness at shorter distances provides echolocation information, (5) at longer distances, timbre aspects, such as sharpness, might be used to detect objects. We also discuss binaural information, movements and the auditory model approach. Autocorrelation was assumed as a proper measure for pitch, but the question is raised whether a mechanism based on strobe integration is a viable possibility.
q-bio.NC
q-bio
1 Human Echolocation in Static Situations: Auditory Models of Detection Thresholds for Distance, Pitch, Loudness and Timbre Bo N. Schenkman1*¶, Vijay Kiran Gidla2¶, 1CTT - Centre for Speech Technology, Department of Speech, Music and Hearing Royal Institute of Technology, Stockholm, Sweden. *Corresponding author, email: [email protected] 2Blekinge Institute of Technology, Karlskrona, Sweden. Email: [email protected] ¶The authors contributed equally to this work. 2 Human echolocation describes how people use reflected sounds to obtain information about their ambient world. We investigated, by using auditory models, how three perceptual parameters, loudness, pitch and sharpness, determine echolocation. We used acoustic recordings from two previous studies, both from stationary situations, and their resulting perceptual data as input to our analysis. An initial analysis was on the room acoustics of the recordings. The parameters of interest were sound pressure level, autocorrelation and spectral centroid. The auditory models were used to analyze echolocation resulting from the perceptual variables, i.e. loudness, pitch and sharpness. Relevant auditory models were chosen to simulate each variable. Based on these results, we calculated psychophysical thresholds for detecting a reflecting object with constant physical size. A non-parametric method was used to determine thresholds for distance, loudness, pitch and sharpness. Difference thresholds were calculated for the psychophysical variables, since a 2-Alternative- Forced-Choice Paradigm had originally been used. We found that (1) blind persons could detect objects at lower loudness values, lower pitch strength, different sharpness values and at further distances than sighted persons, (2) detection thresholds based on repetition pitch, loudness and sharpness varied and depended on room acoustics and type of sound stimuli, (3) repetition pitch was useful for detection at shorter distances and was determined from the peaks in the temporal profile of the autocorrelation function, (4) loudness at shorter distances provides echolocation information, (5) at longer distances, timbre aspects, such as sharpness, might be used to detect objects. We also discuss binaural information, movements and the auditory model approach. Autocorrelation was assumed as a proper measure for pitch, but the question is raised whether a mechanism based on strobe integration is a viable possibility. Keywords: blind; echolocation; psycho-acoustics; auditory models; detection thresholds Author summary Blind people use other senses than vision to orient and move around. Hearing, especially echolocation, i.e. reflection of echoes from objects, might be used to detect objects. Sounds can be produced by the person, by mouth or cane tapping, but can also arise from the environment, e.g. from traffic. We have been studying perceptual processes when blind people use echolocation by reanalyzing results from earlier studies. We used models of hearing to understand how these sounds are processed. Blind people could detect objects in more difficult conditions and at longer distances than sighted people. Detection was based on pitch, loudness and a timbre aspect, sharpness. The perception of pitch was useful at shorter distances and depended on the time properties of the sound and on loudness. At longer distances timbre aspects, such as sharpness, might be used to detect objects. Understanding echolocation is important for how people manage to move around in their surroundings and how our senses process spatial information of the environment. The issues involved are of interest for physiologists, psychologists and biologists. Most importantly, the issues are of practical relevance for blind people learning how to safely move around and for the design of mobility aids. 3 Introduction Persons with blindness use echolocation to obtain information about their surroundings. A person or a source in the environment emits a sound and the reflection is perceived. Both static and dynamic means are used for this sensory information. In both cases the person has to perceive if an object or obstacle is in front of him/her. This perceptual decision is determined by a threshold of detection. The threshold may vary as a function of a number of variables, like the character of the sound emitted, the rates of this sound, its position relative to the person, if motion is involved and the experience and expertise in echolocation. For a review of human echolocation, see Stoffregen and Pittenger [1], Kolarik et al [2] and Thaler and Goodale [3]. Physical properties may have different effects on psychoacoustic parameters that are used to determine if an object is in front or not. Three psychoacoustic parameters are particularly important as sources for human echolocation, viz. pitch in the form of repetition pitch, loudness and spectral information that is perceived as timbre. We describe how the information provided by pitch, loudness or timbre may result in their respective detection thresholds for echolocation. We limit ourselves to stationary situations, i.e. when neither object nor person is moving. When movement is involved, more potential information may be provided (Wilson, [4]; Wallmeier and Wiegrebe, [5]). We also determine at what distance there is a threshold when a person may detect a reflecting object. A number of auditory models were applied to the physical stimuli and we related the results of these models to the perceptual responses of participants from two previous empirical studies, Schenkman and Nilsson [6] and Schenkman, Nilsson and Grbic [7] which also will be referred to as SN2010 and SNG2016, respectively. Psychoacoustic and neuroimaging methods are very useful for describing the high echolocating ability of the blind and their underlying processes. However, they do not fully reveal the information in the acoustic stimulus that determines echolocation (at least when the source for the information is not known) and how this information is encoded in the auditory system. We wanted to know how this information is represented and processed in the human auditory system. One fruitful way to study the information necessary for human echolocation is by signal analysis on the acoustic stimulus. However, such an analysis is only directed at the physical properties of the sound, and does not fully show how the information is represented in the human auditory system. To investigate this, we used auditory models which mimic human hearing. Analyzing the acoustic stimulus using these models provide insight into the processes for human echolocation. They may also allow testing of hypotheses by comparing models (Dau [8]). Loudness, pitch (Schenkman and Nilsson [9]) and timbre are three perceptual attributes of an acoustic sound that are relevant for human echolocation. The backgrounds of these attributes for modeling human echolocation are discussed next. Loudness Loudness is the perceptual attribute of sound intensity and is defined as that attribute of auditory sensation in terms of which sounds can be ordered on a scale from quiet to loud (ASA 1973 [10]). The dynamic range of the auditory system is wide and different mechanisms play a role in intensity discrimination. Psychophysical experiments suggest that neuron firing rates, spread of excitation and phase locking play a role in intensity perception, but the latter two may not always be essential. A disadvantage with the neuron firing rates is that, although the single neurons in the auditory nerve can be used to explain the intensity discrimination, this does not explain why the intensity discrimination is not better than observed, suggesting that the discrimination is limited by the capacity of the higher levels in the auditory system, which may also play a role in intensity discrimination (Moore 2013 4 [11]). Several models (Moore [11] pp 139 - 140) have been proposed to calculate the average loudness that would be perceived by listeners. The basic structure of these models is that, initially the outer and middle ear transformations are performed and then the excitation pattern is calculated. The excitation pattern is transformed into specific loudness, which involves a compressive non-linearity. The total area calculated for the specific loudness pattern is assumed to be proportional to the overall loudness. Therefore, independent of the mechanism underlying the perception of loudness, the excitation pattern is the essential information that is needed for an auditory model of loudness, and thus also for understanding human echolocation in its utilizing of loudness. Pitch Pitch is "that attribute of auditory sensation in terms of which sounds may be ordered on a musical scale" (ASA 1960 [12]). One view of the underlying mechanisms of pitch is that, as the cochlea is assumed to perform a spectrum analysis, the acoustic vibrations are transformed into the spectrum, coded as a profile of discharge rate across the auditory nerve. An alternative view proposes that the cochlea transduces the acoustic vibrations into temporal patterns of neural firing. These two views are known as the place and time hypotheses. According to the place hypothesis, pitch is determined from the position of maximum excitation along the basilar membrane, within the cochlea. This explains how the pitch is perceived by pure tones at low levels, but it fails to explain the perception of pure tones at higher levels. At such levels, due to the non-linearity of the basilar membrane, the peaks become broader and tend to shift towards a lower frequency place. This should lead to a decrease in pitch, but psychophysical experiments show that the pitch is stable. Another case where the place hypothesis fails is its inability to explain the pitch of stimuli whose fundamental is absent. According to the paradox of the missing fundamental, the pitch evoked by a pure tone remains the same if we add additional tones with frequencies that are integer multiples of that of the original pure tone, i.e. harmonics. It also does not change if we then remove the original pure tone, the fundamental (De Cheveigné 2010 [13]). Since the time hypothesis states that pitch is derived from the periodic pattern of the acoustic waveform it overcomes the problem of the missing fundamental. However, the main difficulty with the time hypothesis is that it is not easy to extract one pulse per period, in a way that is reliable and fully general. Psychoacoustic studies also show that pitch exists for sounds which are not periodic. Of interest to our subject matter, human echolocation, is an instance of such sounds, namely iterated ripple noise. It is a sound that models some of the human echolocation signals (e.g. Bilsen [14]). In order to overcome the limitations of the place and time hypothesis two new theories have been proposed, pattern matching (De Boer [15-16]; De Cheveigné [13]), and a theory based on autocorrelation (Licklider [17]; De Cheveigné [13]). De Boer [15-16] described pattern matching such that the fundamental partial is the necessary correlate of pitch, but it may be absent if other parts of the pattern are present. In this way pattern matching supports the place hypothesis. Later Goldstein [18], Wightman [19] and Terhardt [20] described other models for pattern matching. One problem with the pattern matching theory is that it fails to account for pitch whose stimuli have no resolved harmonics. The autocorrelation hypothesis assumes temporal processing in the auditory system. It states that, instead of detecting the peaks at regular intervals, the periodic neural pattern is processed by coincidence detector neurons that calculate the equivalent of an autocorrelation function (Licklider [17]; De Cheveigné [13]). The spike trains are delayed within the brain by various time lags (using neural delay lines) and are combined or correlated with the original. When the lag is equal to the time delay between spikes, then the correlation is high and outputs of the coincidence detectors tuned to that lag are strong. Spike trains in each 5 frequency channel are processed independently and the results are combined into an aggregate pattern. However, De Cheveigné [13] argued that the autocorrelation hypothesis works too well: It predicts that, pitch should be equally salient for stimuli with resolved and unresolved partials, but this is not the case. An alternative to the theory based on an autocorrelation like function is the strobe temporal integration (STI) of Patterson Allerhand, and Giguere [21]. In accordance with STI the auditory image underlying the perception of pitch is obtained by using triggered, quantized, temporal integration, instead of an autocorrelation function. The STI works by finding the strobes from the neural activity pattern and integrating it over a certain period. Thus, there is no full understanding of how pitch is perceived. Irrespective if temporal, spectral or multi mechanisms determine pitch perception, the underlying information that the auditory system uses to detect pitch has to be the excitation pattern on the basilar membrane. Hence, the excitation pattern is the crucial information that should be simulated by an auditory model for pitch perception, and thus also for human echolocation. Repetition pitch Human echolocation signals consist of an original sound along with a reflected or delayed signal. Several studies have been presented to explain the pitch perception of such sounds. Bassett and Eastmond [22] examined the physical variations in the sound field close to a reflecting wall. They reported a perceived pitch caused by the interference of direct and reflected sound at different distances from the wall; the pitch value being equal to the inverse of the delay. In a similar way, Small and McClellan [23] and Bilsen [24], delayed identical pulses and found that the pitch perceived was equal to the inverse of the delay, naming it time separation pitch, and repetition pitch, respectively. When a sound and the repetition of that sound are listened to, a subjective tone is perceived with a pitch corresponding to the reciprocal value of the delay time (Bilsen and Ritsma [25]). They explained the repetition pitch phenomenon by the autocorrelation peaks or the spectral peaks. Yost [26] performed experiments using iterated ripple noise stimuli and concluded that autocorrelation was the underlying mechanism, used by the listeners to detect repetition pitch. Timbre When the loudness and pitch of an acoustic sound are similar, the subjective attribute of sound which may distinguish or identify the sound is its timbre. Timbre has been defined as that attribute of an auditory sensation which enables a listener to judge that two non-identical sounds, similarly presented and having the same loudness and pitch, are dissimilar (ANSI 1994 [27]). One example is the difference between two musical instruments playing the same tone, e.g. a guitar and a piano. Timbre is a multidimensional percept and there is no single scale on which we can order timbre. To quantify timbre one approach is to consider the overall distribution of the spectral energy. Plomp and co-workers [28-29] showed that the perceptual differences between different sounds were closely related to the levels in 18 1/3 octave bands, thus relating the timbre to the relative level produced by the sound in each critical band. Hence, generally, for both speech and non-speech sounds, the timbre of steady tones is determined by their magnitude spectra, although the relative phases may play a small role [30]. When we consider time varying patterns, there are several factors that may influence the perception of timbre: (i) periodicity; (ii) variation of the envelope of the waveform; (iii) spectrum changes over time; and (iv) what the preceding and following sounds were like. Using auditory models, timbre information can be assessed by the levels in the spectral envelope and by the variations of the temporal envelope. Another way to preserve the fine grain time interval information that is necessary for timbre perception is by the strobe temporal integration (STI) method of Patterson, Allerhand, and Giguere [21]. 6 and PsySound3 were Signal analysis Signal analysis and auditory models may enable us to understand the processing of sounds for persons using echolocation, since one then can consider the transmission of the acoustic sound from the source via the internal representation to the final percept of the person. The acoustic sound travels and undergoes transformation because of the room acoustics. One should therefore first understand the information that is received at the human ear. Signal analysis is useful for this purpose, as we then can analyze the characteristics of the sound, which have been transformed due to room conditions. The second step is to analyze how the desired characteristics of the acoustic sound, that contains the information, are represented in the auditory system. Here auditory models are useful. The desired information is transformed in an analogous way to how the auditory system is known to process it. Keeping track of the information from the outer ear to the central nervous system will be an important part for describing how listeners perceive sounds and explaining the differences between groups of listeners with different characteristics, e.g. visually handicapped vs sighted persons. This is the methodology which was used for this report. downloaded To model the auditory analysis performed by the human auditory system we used the auditory image model of Patterson, Allerhand, and Giguere [21], the loudness models of Glasberg and Moore [31-32] and the sharpness model of Fastl and Zwicker [33]. Matlab was used as the implementation environment. The auditory image model has been implemented in matlab by Bleeck, Ives, and Patterson [34] and the current version is known as AIM-MAT. The loudness and the sharpness models were implemented in PsySound3 (Cabrera, Ferguson, and Schubert [35]), a GUIdriven Matlab environment for analysis of audio recordings. AIM- MAT from https://code.soundsoftware.ac.uk/projects/aimmat and http:// www.psysound.org, respectively. Aims and hypothesis The aims of the present study were: (1) To study different components of the information in the acoustic stimulus, that determines echolocation. (2) To determine the thresholds for different components of the information in the acoustic stimulus, that are important factors for the detection distance to reflecting objects. (3) To find out how the acoustic information that determines the high echolocation ability of the blind is represented in the human auditory system. More specifically, our hypotheses were: (1) Detection thresholds based on repetition pitch, loudness and sharpness will vary and will depend on the room acoustics and type of the sound stimuli that is used. (2) Repetition pitch is useful for detection at shorter distances and is determined from the peaks in the temporal profile of the autocorrelation function, computed on the neural activity pattern. (3) Detection at shorter distances, based on loudness provides information for listeners. (4) At longer distances timbre aspects, such as sharpness information might be used by listeners to detect objects. Structure of the report This report is hereafter structured as follows: In the next part "Method", subtitled "Room acoustics: Basic acoustic analyses of physical parameters" we describe the recordings used in the studies by Schenkman and Nilsson, SN2010 [6] and Schenkman, Nilsson, and Grbic SNG2016 [7], since these form the data for the present report. In the same part we present a 7 signal analysis conducted on these recordings. We describe basic room acoustic parameters of the signals that form the basis for the physical information on the room to the persons, whether objects are present or not. No consideration is here done to auditory models. In the next part, "Models", subtitled "Auditory analysis of acoustic information", we describe the auditory models, their designs and implementation. The loudness, pitch and sharpness for the recordings of SN2010 [6] and SNG2016 [7] when analyzed using the auditory models are presented in the Results, "Loudness analysis: Excitation patterns, binaural loudness, short and long term loudness", "Pitch analysis: autocorrelation with dual profiles" and "Sharpness analysis" sections, respectively. In the section thereafter, "Threshold values, absolute and difference, for echolocation in static situations based on auditory model analysis," the thresholds for object detection are presented. This is followed by Discussion and Conclusions. Method Room acoustics: Basic acoustic analyses of physical parameters Sound recordings used Here we describe briefly how the sound recordings of SN2010 [6] and SNG2016 [7] were made. For more detailed descriptions, see the original articles. In SN2010 [6], the binaural sound recordings were conducted in an ordinary conference room and in an anechoic chamber using an artificial manikin. The object was a reflecting 1.5 mm thick aluminum disk with a diameter of 0.5 m. Recordings were conducted at distances of 0.5, 1, 2, 3, 4, and 5 m between microphones and the reflecting object. In addition, recordings were made with no obstacle in front of the artificial manikin. Durations of the noise signal were 500, 50, and 5 ms; the shortest corresponds perceptually to a click. The electrical signal was a white noise. However, the emitted sound was not perfectly white, because of the non-linear frequency response of the loudspeaker and the system. A loudspeaker generated the sounds, resting on the chest of the artificial manikin. In SNG2016 [7] recordings were conducted in an ordinary lecture room. Recordings were conducted at 100 and 150 cm distances between microphones and the reflecting object. The emitted sounds were either bursts of 5 ms each, varying in rates from 1 to 64 bursts per 500 ms or a 500 ms white noise. In contrast to SN2010 [6], the sounds in SNG2016 [7] were generated by a loudspeaker placed 1 m straight behind the center of the head of the artificial manikin. The sound recording set ups can be seen in Fig 1. 8 Fig 1. Sound recordings used. (A) anechoic room (B) conference room, with loudspeaker on the chest of the artificial manikin in Schenkman and Nilsson [6] (C) lecture room with loudspeaker behind the artificial manikin in Schenkman, Nilsson, and Grbic [7]. Calibration of signals Analyses were performed to determine basic room acoustical parameters relevant for human echolocation: sound pressure level, autocorrelation, and spectral centroid. Before analyzing the recordings, the recordings were calibrated by the calibrating constants (CC) using Equation (1). Based on the Sound Pressure Level (SPL) of 77, 79 and 79 dBA for the 500 ms recording without the object at the ear of the artificial manikin in the anechoic and conference rooms of SN2010 and in the lecture room of SNG2016, the CC's were calculated to be 2.4663, 2.6283 and 3.5021, respectively. A-weighting was not included in Equation (1), since the difference of using it was less than 0.5 dB and could thus be neglected. See S.1 Supporting Information for more details. 𝐶𝐶𝐶𝐶= 10�𝑆𝑆𝑆𝑆𝑆𝑆−20∗𝑙𝑙𝑙𝑙𝑙𝑙10�𝑟𝑟𝑟𝑟𝑟𝑟(𝑟𝑟𝑠𝑠𝑙𝑙𝑠𝑠𝑠𝑠𝑙𝑙) 20∗10−6 � 20 � (1) As the recordings were binaural, both left and right ear recordings were analyzed. The recordings in SN2010 [6] and SNG2016 [7] had 10 versions of each duration and distance. For the condition with no reflecting object, there were two sets of recordings in SN2010 [6], 10 versions in each, while for the same condition in SNG2016 [7] there was one set of recordings with 10 versions. The two versions with no object in SNG2010 [6], resulted 9 in very similar values, but are for fullness presented separately in this report. It should be noted that the recordings vary over the versions causing the term "rms(signal)" in Equation (1) to vary, thereby varying the calibrated constants for the 10 different versions. However, as the variation was very small between the versions, we decided to use only the 9th version of the 500 ms first recording without the object in SN2010 [6] and the 9th version of the 500 ms recording without the object in SNG2016 [7] to establish the calibrated constants. Another reason to choose only one version, the 9th, is that although the other versions may not have exactly the identical CC's they will be relatively calibrated with respect to the recording of version 9. For example, suppose the recording in the anechoic chamber version 1 had 67 dB SPL and version 9 had 66 dB SPL before calibration, then the levels obtained by calibrating the recordings to 77 dB SPL using the CC of the 9th version would be 78 dB SPL for version 1 and 77 dB SPL for version 9. In other words, they will give the same level difference, also after calibration. Sound Pressure Level (SPL) Detection of objects by echolocation is to a certain extent based on intensity information. Hence, the SPL in dBA were calculated using Equation (2), where "RMS" is the root mean square amplitude of the signal analyzed. As has been pointed out by various authors (e.g. Rowan et al. [36]), binaural information may be utilized for echolocation purposes. We therefore calculated the SPL values for both ears. The mean SPL values of the 500 ms recordings in study SN2010 [6] and in study SNG2016 [7] are shown in Tables 1 and 2, respectively. The values for the 5 ms and 50 ms recordings are for reasons of space not shown here. As mentioned above, for the recordings with no object, two series were conducted in SN2010, each with 10 recordings. As can be seen in Table 1, the values are very close to each other. 𝑆𝑆𝑆𝑆𝑆𝑆=20∗𝑙𝑙𝑙𝑙𝑙𝑙10�𝐶𝐶𝐶𝐶∗𝑟𝑟𝑟𝑟𝑟𝑟(𝑟𝑟𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠) 20∗10−6 � (2) Table 1. Mean of the sound pressure levels (dBA) for the left and right ears over the 10 versions of the 500 ms duration signals in the anechoic and conference room used by Schenkman and Nilsson [6]. For the recording with no object, two series were conducted. Object Distance (cm) No Object, recording 1 No Object, recording 2 50 100 200 300 400 500 Anechoic chamber Conference room Left ear Right ear Left ear Right ear 77.2 77.6 85.2 81.9 77.1 77 77.1 77 77.9 77.4 88.2 82.6 78 78.2 78 78 79 79 87.5 82.8 79.6 78.9 79 79 78.8 78.8 87.5 82.4 79.5 78.9 78.9 78.8 10 Table 2. Mean of the sound pressure level (dBA) for the left and right ears over the 10 versions of the 500 ms duration signals in the lecture room used by Schenkman, Nilsson and Grbic [7]. Lecture room Object Distance (cm) Left ear Right ear No Object 100 150 79.2 79.6 79.4 79.6 81.5 79.7 The tabulated SPL values in Tables 1 and 2 show the effect of room acoustics on level differences, between the ears and between the rooms. The level differences between the recording without object and the recordings with object were in SNG2016 less than those in SN2010. This may be due to the differences in experimental setup (Fig 1) or to the acoustics of the room. The extent to which this information affected the listeners in these studies is not obvious, as loudness perceived by the human auditory system cannot be related directly to the SPL (e.g. Moore [11]). This issue is analyzed further in the section "Loudness analysis: Excitation patterns, binaural loudness, short and long term loudness". Autocorrelation Function (ACF) Repetition pitch is an important aspect of how we perceive complex sounds (Bilsen [24]; Bilsen and Ritsma [25]). Schenkman and Nilsson [9] showed that this pitch, rather than loudness, is used by listeners to detect an object by echolocation. As noted above, pitch perception can often be explained by the peaks in the autocorrelation function and therefore an autocorrelation analysis was performed, which we present here. The theoretical values for repetition pitch for the recordings of SN2010 [6] and SNG2016 [7] were calculated using Equation (3). The corresponding values for recordings with objects at distances of 50, 100, 150, 200, 300, 400 and 500 cm would be approximately 344, 172, 114, 86, 57, 43 and 34.4 Hz, assuming sound velocity to be 344 m/s. As the theory based on autocorrelation uses temporal information, repetition pitch perceived at the above frequencies can be explained by the peaks in the ACF at the inverse of the frequencies, i.e. approximately at 2.9, 5.8, 8.7, 11.6, 17.4, 23.2 and 29 ms, respectively. The autocorrelation analysis was performed using a 32 ms frame, which would cover the required pitch period. A 32 ms hop size was used to analyze the ACF for the next time instants of 64 ms, 96 ms etc. In order to compare the peaks among all the recordings, the ACF was not normalized to the limits -1 to 1. 𝑅𝑅𝑆𝑆= 𝑟𝑟𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 𝑜𝑜𝑜𝑜 𝑟𝑟𝑜𝑜𝑠𝑠𝑠𝑠𝑠𝑠 2∗𝑠𝑠𝑠𝑠𝑟𝑟𝑑𝑑𝑠𝑠𝑠𝑠𝑑𝑑𝑠𝑠 𝑜𝑜𝑜𝑜 𝑑𝑑ℎ𝑠𝑠 𝑜𝑜𝑜𝑜𝑜𝑜𝑠𝑠𝑑𝑑𝑑𝑑 where RP is Repetition Pitch. (3) In the study SN2010 [6] the participants performed well with the longer duration signals. For a single short burst the person had only one chance to perceive the signal and its echo. This can be visualized from the ACFs in Figs 2 and 3, where for the 5ms recording the peak was present only for the initial 32 ms frame. For the 500 ms recording the peak was also present for frames with time instants greater than 32 ms. (Note that for each duration of the signals an additional 450 ms silence was padded and presented to the test persons. The ACF were analyzed in the same manner, and hence the 5ms duration signal had a total duration of 455 ms and the 500 ms signal had a total duration of 950 ms.) 11 Fig 2. The autocorrelation function of a 5 ms signal recorded in the anechoic chamber (SN2010) [6] with reflecting object at 100 cm. The sub figures A-E show the autocorrelation function (ACF index) at 32, 64, 96, 128, 160 and 192 ms time instants of the signal (Lag), respectively. As the recording is only 5ms in duration the autocorrelation function is only present in the first 32ms frame. Fig 3. The autocorrelation function of a 500 ms signal recorded in the anechoic chamber (SN2010) [6] with reflecting object at 100 cm. The sub figures A-E show the autocorrelation function (ACF index) at 32, 64, 96, 128, 160 and 192 ms time instants of the signal (Lag), respectively. Schenkman and Nilsson [6] argued that the higher detection ability of their participants for the longer duration signals may have been a result of attention, a confounding 12 factor. Even if a test person at first was not attentive, he or she had a longer time interval available to perceive the signal. However, in Schenkman, Nilsson and Grbic [7] the performance decreased at one distance for the longer duration noise, although the repetitions were present for the frames with a time instant greater than 32 ms. These ACF functions are for reasons of space not shown here. Therefore, that longer duration signals are always beneficial for human echolocation cannot be concluded on the available results. The peak heights at the pitch period for the recordings with object at 100 cm for the 5 ms duration signal in the conference room in SG2010 [6] were greater than those in the lecture room in SNG2016 [7]. The 500 ms duration signal with object at 100 cm in the lecture room in SN2016 [7] had a greater peak height than the 5 ms signal in the conference room in SN2010 [6], but the peak is not distinct enough when compared to the 500 ms duration signal in the conference room in SN2010 [6]. The cause for these differences in the peak heights between the two rooms, conference room in SN2010 [6] and the lecture room in SNG2016 [7], are probably the different room acoustics. The ACF depends on the spectrum of the signal, and the acoustics of the room certainly influences the peaks in the ACF. The reverberation time, T60, for the conference and the lecture room were 0.4 and 0.6 seconds, respectively. A fuller discussion of how the information carried by the peaks is represented in the auditory system is further discussed in the section "Pitch analysis: autocorrelation with dual profiles". Timbre: Spectral Centroid (SC) Detection of an object could be provided by the timbre information available in a sound. Timbre is constituted by a number of different characteristics, e.g. roughness. Another characteristic of timbre perception is the spectral centroid (Peeters et al. [37]), which gives a time varying value characterizing the subjective center of the timbre for a sound. As mentioned in the Introduction, in the timbre section previously, the timbre of steady tones are mainly determined by their magnitude spectra. We believe that the spectral centroid is an important feature of human echolocation in static situations. As the recordings in SN2010 [6] and SN2016 [7] were static, the spectral centroid using the magnitude spectra of the recordings was computed to depict the timbre information. To compute the spectral centroid, the recordings were analyzed using a 32 ms frame with a 2 ms overlap. The spectral centroid for each frame was computed by Equation (4). As the spectral centroid for each frame is a time varying function, it is plotted as a function of time. The means of the spectral centroid for the 10 versions at each condition for the 500 ms of the left ear recordings are shown in Figs 4 to 6. 𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑙𝑙𝐶𝐶𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆= ∑(𝐹𝐹𝑟𝑟𝑠𝑠𝐹𝐹𝑠𝑠𝑠𝑠𝑠𝑠𝑑𝑑𝐹𝐹∗𝐹𝐹𝐹𝐹𝐹𝐹(𝑜𝑜𝑟𝑟𝑠𝑠𝑟𝑟𝑠𝑠)) ∑(𝐹𝐹𝐹𝐹𝐹𝐹(𝑜𝑜𝑟𝑟𝑠𝑠𝑟𝑟𝑠𝑠)) In SN2010 [6] for the recordings without the object, the spectral centroid was approximately below 5000 Hz. For the recordings with the object at 50 and 100 cm, the spectral centroids were approximately above 5000 Hz (Fig 4). This difference might provide information to listeners to distinguish conditions with an object from those without an object. The recordings with the object at 200 to 500 cm did not vary much when compared with the recording without the object. In SNG2016 [7] the spectral centroid was approximately 6000 Hz for all recordings (Fig 6), showing very small changes. The timbre information in SNG2016 [7] may thus at first sight not seem to be useful for echolocation. This physical analysis indicates that there was variation in the spectral centroid in the recordings of SN2010 [6] with object at shorter distances (distances shorter than 200 cm) but for longer distances the difference in the spectral centroid was almost negligible. (4) 13 Fig 4. The means of the spectral centroid of the 10 versions as a function of time of the left ear for a 500ms recording in the anechoic chamber (SN2010) [6]. The sub figures A, B are for the two no object recordings and C-H are for the recordings with object at 50, 100, 200, 300, 400 and 500 cm, respectively. Fig 5. The means of the spectral centroid of the 10 versions as a function of time of the left ear for a 500 ms recording in the conference room (SN2010) [6]. The sub figures A, B are for the two no object recordings and C-H are for the recordings with object at 50, 100, 200, 300, 400 and 500 cm, respectively. The conclusions above are based on a purely physical analysis, a Fast Fourier Transform (FFT) analysis of the sounds. However, the spectral analysis performed by the auditory system is more complex than a FFT that we have used to compute the spectral 14 centroid. In the next section we will show that the conclusions will be modified, when we consider how human hearing works. This we do by using auditory models to analyze the sounds. Fig 6. The mean of the spectral centroid for the 10 versions as a function of time of the left ear for a 500 ms recording in the lecture room (SNG2016) [7]. The sub figure A is for the no object recording, while sub figures B and C are for the recordings with object at 100 and 150 cm, respectively. Models Auditory analysis of acoustic information Premises In the previous sections in the Method part, we presented physical parameters from the two studies SN2010 [6] and SNG2016 [7], that determine human echolocation. In this Models part we take the physical parameters from above and study how they may provide the basis for relevant auditory information to a person. In the Results part thereafter, we will connect this auditory analysis with the behavioural results. For the auditory analysis we used the auditory image model (AIM), originally developed by Patterson et al. [38, 21] with extensions added by other authors. It is a time- domain, functional model of the signal processing performed in the auditory pathway as the system converts a sound wave into the perception that we experience when presented with a sound. This representation is referred to as an auditory image by analogy with the visual image of a scene that we experience in response to optical stimulation. The AIM simplifies the peripheral and the central auditory systems into modules. A summarily description of the AIM and how the modules were implemented in the present analysis is given below. A more detailed description of each module of AIM can be found at http://www.acousticscale. org/wiki/index.php/AIM2006_Documentation. We used the modules described below to analyse the recordings. All the processing modules of AIM are written in matlab. The current version, as mentioned earlier, is referred to as AIM-MAT and can be downloaded from https://code.soundsoftware.ac.uk/projects/aimmat. The autocorr module was only present in the 2003 version of AIM and can be downloaded from http://w3.pdn.cam.ac. uk/groups/cnbh/aimmanual/download/downloadframeset.htm. 15 First step: Pre Cochlear Processing (PCP) The outer middle ear transformation of the acoustic sound is simulated in AIM by the Pre Cochlear Processing (PCP) module. The PCP module consists of four different Finite Impulse Response (FIR) filters, designed for different applications. These are: (i) Minimum audible field (MAF), which is suitable for signals presented in free field. (ii) Minimum audible pressure (MAP), which is suitable for systems which produce a flat frequency response. (iii) Equal loudness contour (ELC) and (iv) filter gm2002 (Glasberg and Moore [31]) are almost identical and include the factors associated with the extra internal noise at low and high frequencies. However, gm2002 uses more recent data of Glasberg and Moore [31]). The MAF, MAP, ELC were designed using Parks-McClellan optimal equi-ripple FIR filter design algorithm, while the gm2002 was designed using a frequency sampling method. The transmission of the acoustic sound through the PCP filter can be modelled using Equation (5), where Signalinput is the input to the AIM and Signalpcp is the filtered output of the corresponding PCP filter. An example of the frequency response used to generate a PCP filter is shown in Fig 7. 𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑠𝑠𝑑𝑑𝑠𝑠=𝑓𝑓𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑆𝑆�𝑆𝑆𝐶𝐶𝑆𝑆𝑜𝑜𝑠𝑠𝑠𝑠𝑑𝑑𝑠𝑠𝑟𝑟,𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑑𝑑� (5) Fig 7. The frequency response used to design the gm2002 filter of the Pre Cochlear Processing (PCP) module in the AIM. The frequency response was obtained from the frontal field to cochlea correction data of Glasberg and Moore [31]. Second step: Basilar Membrane Motion (BMM) The non-linear spectral response of the basilar membrane is an important feature of the peripheral auditory system. This response is implemented in the AIM by a dynamic compressive gammachirp filter bank, dcGC (Irino and Patterson [39]). Two important properties of the Basilar Membrane Motion (BMM) are the asymmetry and the compression of the auditory filters that are made in proportion to the intensity level. These properties were designed using a compressive gammachirp filter (cGC). It is a generalized form of the gammatone filter, which was derived with operator techniques (Irino and Patterson [39]). The developments of both the gammatone and gammachirp filters are described in Patterson, Unoki, and Irino [40]. The cGC is simulated by cascading a passive gammachirp filter (pGC) with a high pass asymmetric function (HP-AF). The asymmetrical property is simulated by 16 the pGC filter and its output is used to adjust the level dependency of the active part, i.e. the HP-AF. Other options available for generating the BMM in AIM are the gammatone function and the pole zero filter cascade. Since the gammatone function does not depict the non- linearity of the basilar membrane, we used the default filterbank dcGC to simulate the BMM. The transformation of the BMM can be modelled using Equations (6) and (7). SignalpGC(fc) is the filtered output of the pGC filterbank, fc is the centre frequency of the filter, ACF(fc) is the high pass asymmetric compensation filters and SignalcGC(fc) is the final compressed output of the BMM. For a detailed description of the pGC and cGC filterbanks, see Irino and Patterson [39]. 𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑠𝑠𝑝𝑝𝐶𝐶(𝑓𝑓𝑑𝑑)=𝑓𝑓𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑆𝑆�𝑆𝑆𝑃𝑃𝐶𝐶(𝑓𝑓𝑑𝑑),𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑠𝑠𝑑𝑑𝑠𝑠� (6) 𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑑𝑑𝑝𝑝𝐶𝐶(𝑓𝑓𝑑𝑑)=𝑓𝑓𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑆𝑆�𝐴𝐴𝐶𝐶𝐴𝐴(𝑓𝑓𝑑𝑑),𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑠𝑠𝑝𝑝𝐶𝐶(𝑓𝑓𝑑𝑑)� (7) Third step: Neural Activity Pattern (NAP) The basilar membrane motion is transduced into an electrical potential by the inner hair cells. The Neural Activity Pattern (NAP) is implemented in AIM by half wave rectification followed by low pass filtering. Low pass filtering is executed as phase locking is not feasible for high frequencies in the human ear. There are three modules in the AIM to generate the NAP: (i) half wave rectification followed by compression and low pass filtering (H-C-L) (ii) half wave rectification followed by low pass filtering (H-L) (iii) two dimensional adaptive threshold (similar to H-C-L but it has adaptation which is more realistic). The choice of NAP module depends on the choice of BMM module. As noted above, we used a dcGC filter bank in our analyses, and the compression of the basilar membrane was simulated by it. The H-L module was therefore chosen to generate the NAP. This transformation can be modelled using Equation (8), where abs(Signalbmm(fc)) is the half wave rectified signal of the basilar membrane, fc is the centre frequency of the filter, LPF is the low pass filter and Signalnap(fc) is the modelled NAP. 𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑠𝑠𝑠𝑠𝑠𝑠(𝑓𝑓𝑑𝑑)=𝑓𝑓𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑆𝑆(𝑆𝑆𝑆𝑆𝐴𝐴,𝑆𝑆𝑎𝑎𝑎𝑎(𝑆𝑆𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑙𝑙𝑜𝑜𝑟𝑟𝑟𝑟(𝑓𝑓𝑑𝑑))) (8) Fourth step: Strobe Temporal Integration (STI) The fourth stage in the AIM represents processing in the central nervous system. Perceptual research suggests that at least some of the fine grain time interval information is needed to preserve timbre information (Krumbholz et al [41], Patterson [42 - 43]). Auditory models often time average the NAP information, which unfortunately then loses the fine grain information. To prevent this, AIM uses a procedure called Strobe Temporal Integration (STI), which is subdivided into two modules, (i) strobe finding, and (ii) temporal integration. Strobe Finding (SF): The sub module sf2003 is used to find the strobes from the NAP. It uses an adaptive strobe threshold to issue a strobe and the time of the strobe is that associated with the peak of the NAP pulse. After the strobe is initiated the threshold initially rises along a parabolic path and then returns to the linear decay to avoid spurious strobes. The duration of the parabola is proportional to the centre frequency of the channel and the height to the height of the strobe. After the parabolic section of the adaptive threshold, its level decreases linearly to zero in 30 ms. An additional feature of sf2003 is the inter channel interaction, i.e. a strobe in one channel reduces the threshold in the neighbouring channels. An example of how the threshold varies and how the strobes are calculated is shown in Fig 8. 17 Fig 8. The Neural Activity Pattern (NAP) of a 200 Hz pure tone in the 253 Hz frequency channel. The dashed line shows the threshold variation and the dots indicate the calculated strobes. Temporal Integration (TI): The temporal integration is implemented in AIM by a module called stabilized auditory image (SAI). The SAI in its turn uses a sub module, ti2003, to accomplish this. The ti2003 module changes the time dimension of the NAP into a time interval dimension. This works as follows: Initially, a temporal integration is initiated when a strobe is detected. If no further strobes are detected, the process continues for 35 ms and stops. If strobes are detected within the 35 ms interval, each strobe initiates a temporal integration process. To preserve the shape of the SAI to that of the NAP, ti2003 uses weighting. The new strobes are initially weighted high (the weights are also normalized so that the sum of the weights is equal to 1) making the older strobes contribute relatively less to the SAI. Autocorrelation Function (ACF) As was mentioned above, the AIM offers a module, autocorr, to analyze autocorrelation processes. Corresponding physiological processes are presumed to take place in the central nervous system (e.g. Licklider [17]; Bilsen [24; Yost [26]). By using the autocorr module one can implement models of hearing based on autocorrelation processes. The autocorr module takes the NAP as input and computes the ACF on each center frequency channel of the NAP by using a duration of 70 ms, hop time of 10ms and a maximum delay of 35ms. Results Loudness analysis: Excitation patterns, binaural loudness, short and long term loudness In the room acoustics part earlier, for the sound pressure level analysis we described the physical intensity of a sound that may affect human echolocation. This description is necessary for understanding echolocation, but it is not sufficient. Intensity is related to the perception of loudness which is a psychological attribute, but loudness also depends on a number of other parameters, primarily but not only, frequency selectivity, bandwidth and duration of the sound. In this section we consider perceptual aspects of loudness for echolocating sounds, using the loudness model of Glasberg and Moore [31]. We chose this model instead of the AIM for the following reason. 18 The loudness model of Glasberg and Moore (2002) computes the frequency selectivity and compression of the basilar membrane in two stages, (1) by computing the excitation pattern and (2) by the specific loudness of the input signal. Physiologically they are interlinked and a time domain filter bank which simulates both the selectivity and the compression might be appropriate. Although there are different time domain models of the level dependent auditory filters available in AIM (e.g. dcGC), they do not give a sufficiently good fit to the equal loudness contours in ISO 2006 (Moore [44]). Since we consider this fit to be an important aspect, so instead of choosing the AIM model to model loudness, we used the model of Glasberg and Moore [31]. A loudness model should consider the outer middle ear filtering, the non-linearity of the basilar membrane and the temporal integration of the auditory system. The loudness model of Glasberg and Moore [31] estimates the loudness of steady sounds and of time varying sounds, by accounting for these features of the human auditory system. Each stage of this model is described briefly below. Outer middle ear transformation The outer middle ear transformation was modelled using a finite impulse response (FIR) filter with 4097 coefficients. The response at the inner ear can be represented using Equation (9), where x and yomt are the signals before and after transformation, and h is the impulse response of the filter. 𝑦𝑦𝑜𝑜𝑟𝑟𝑑𝑑=𝑓𝑓𝑆𝑆𝑙𝑙𝑆𝑆𝑆𝑆𝑆𝑆(ℎ,𝑥𝑥) (9) Excitation pattern The excitation pattern is defined as the magnitude of the output of each auditory filter as a function of the filter center frequency. To compute the excitation pattern from the time domain signal, Glasberg and Moore [31] used six FFTs in parallel based on Hanning- windowed segments with durations of 2, 4, 8, 16, 32 and 64 ms, all aligned at their temporal centers. The windowed segments are zero padded, and all FFTs are based on 2048 sample points. All FFTs are updated at 1 ms intervals and each FFT was used to calculate the spectral magnitudes at specific frequency ranges. Values outside the range were discarded. The running spectrum was the input to the auditory filters, and their output was calculated at the center frequency of 0.25 Equivalent Rectangular Bandwidth (ERB) intervals taking into account the known variation of the auditory filter shape regarding center frequency and level. The excitation pattern is defined as the output of the auditory filter as a function of center frequency (Glasberg and Moore [31]). This can be represented with Equation (10), where E(fc) is the magnitude of the output of each auditory filter with center frequency fc, Yomt is the power spectrum of yomt calculated using six parallel FFT's, as mentioned above, over a 1ms interval and W(fc) is the frequency response of the auditory filter at center frequency fc . 𝐸𝐸(𝑓𝑓𝑑𝑑)= 𝑌𝑌𝑜𝑜𝑟𝑟𝑑𝑑∗𝑊𝑊(𝑓𝑓𝑑𝑑) (10) Specific loudness (SL) To model the non-linearity of the basilar membrane, the excitation pattern has to be converted to specific loudness. Specific loudness is the loudness in a critical band. This conversion is done in the model of Glasberg and Moore [31] using three conditions (Equation 11). 19 𝑆𝑆𝑆𝑆(𝑓𝑓𝑑𝑑)=⎩⎪⎨⎪⎧𝐶𝐶∗� 𝐸𝐸(𝑜𝑜𝑐𝑐)+ 𝐹𝐹𝑄𝑄(𝑜𝑜𝑐𝑐)�1.5+((𝑃𝑃∗𝐸𝐸(𝑓𝑓𝑑𝑑)+𝐴𝐴 )𝛼𝛼− 𝐴𝐴𝛼𝛼) 2𝐸𝐸(𝑜𝑜𝑐𝑐) ((𝑃𝑃∗𝐸𝐸(𝑓𝑓𝑑𝑑)+𝐴𝐴 )𝛼𝛼− 𝐴𝐴𝛼𝛼) 𝐶𝐶∗� 𝐸𝐸(𝑜𝑜𝑐𝑐) 1.04∗106�0.5 𝑠𝑠𝑜𝑜 𝐸𝐸(𝑜𝑜𝑐𝑐)≤ 𝐹𝐹𝑄𝑄(𝑜𝑜𝑐𝑐) 𝑠𝑠𝑜𝑜 1010 ≥𝐸𝐸(𝑜𝑜𝑐𝑐)≥ 𝐹𝐹𝑄𝑄(𝑜𝑜𝑐𝑐) 𝑠𝑠𝑜𝑜 𝐸𝐸(𝑜𝑜𝑐𝑐) ≥ 1010 (11) TQ(fc) is the threshold of excitation, which is frequency dependent. G represents the low level gain in the cochlear amplifier, relative to the gain at 500 Hz and above, and is frequency dependent. The parameter A is used to bring the input-output function close to linear around the absolute threshold. α is a compressive exponent which varies between 0.20 and 0.27. C is a constant which scales the loudness to conform to the sone scale, where the loudness of 1 kHz tone at 40 dB SPL corresponds to 1 sone and C is equal to 0.047. Loudness depends on the intensity and bandwidth of the sound, but among other factors, also on its duration. Duration of signals is of relevance for human echolocation, and we will therefore briefly describe models of duration of loudness. The effect of the duration on the loudness was modeled by Glasberg and Moore [31] using three concepts for duration of sounds, viz. Instantaneous loudness, Short Term Loudness and Long Term Loudness. They depict the temporal integration of loudness in the auditory system and are described next. Instantaneous loudness (IL). The specific loudness in each critical band has a pattern and the specific loudness over all critical bands is called specific loudness pattern. Usually the area under the specific loudness pattern is summed to give the instantaneous loudness. If the sound is binaural, then the area under the specific loudness patterns at the two ears are summed together to give the instantaneous loudness. The instantaneous loudness is an intervening variable which is used for calculations and is not a perceptual variable. Short Term Loudness (STL). The Short Term loudness is determined by averaging the instantaneous loudness using an attack constant, αa = 0.045, and a decay constant, αr = 0.02 (Equation 12). The values of αa and αr were chosen so that the model will give reasonable predictions for variations of loudness with duration and amplitude modulated sounds (see Moore [44]). 𝑆𝑆𝑆𝑆𝑆𝑆 (𝑆𝑆)= �∝𝑠𝑠∗ 𝐼𝐼𝑆𝑆𝑠𝑠+(1− ∝𝑠𝑠)∗ 𝑆𝑆𝑆𝑆𝑆𝑆𝑠𝑠−1 ∝𝑟𝑟∗ 𝐼𝐼𝑆𝑆𝑠𝑠+(1− ∝𝑟𝑟)∗ 𝑆𝑆𝑆𝑆𝑆𝑆𝑠𝑠−1 𝑠𝑠𝑜𝑜 𝐼𝐼𝐼𝐼(𝑠𝑠) ≥ 𝑆𝑆𝐹𝐹𝐼𝐼(𝑠𝑠−1) 𝑠𝑠𝑜𝑜 𝐼𝐼𝐼𝐼(𝑠𝑠) ≤ 𝑆𝑆𝐹𝐹𝐼𝐼(𝑠𝑠−1) Long Term Loudness (LTL). The Long Term Loudness parameter was calculated by averaging the instantaneous loudness using an attack constant, αa1 = 0.01 and a decay constant, αr1 = 0.0005 (Equation 13). The values of αa1 and αr1 were chosen so that the model may give reasonable predictions for the overall loudness of sounds that are amplitude modulated at low rates (Moore [44]). 𝑆𝑆𝑆𝑆𝑆𝑆 (𝑆𝑆)= �∝𝑠𝑠1∗ 𝑆𝑆𝑆𝑆𝑆𝑆𝑠𝑠+(1− ∝𝑠𝑠1)∗ 𝑆𝑆𝑆𝑆𝑆𝑆𝑠𝑠−1 ∝𝑟𝑟1∗ 𝑆𝑆𝑆𝑆𝑆𝑆𝑠𝑠+(1− ∝𝑟𝑟1)∗ 𝑆𝑆𝑆𝑆𝑆𝑆𝑠𝑠−1 𝑠𝑠𝑜𝑜 𝑆𝑆𝐹𝐹𝐼𝐼(𝑠𝑠) ≥ 𝐼𝐼𝐹𝐹𝐼𝐼(𝑠𝑠−1) 𝑠𝑠𝑜𝑜 𝑆𝑆𝐹𝐹𝐼𝐼(𝑠𝑠) ≤ 𝐼𝐼𝐹𝐹𝐼𝐼(𝑠𝑠−1) As noted above, loudness is also affected by binaural hearing. To model binaural loudness, a number of psychoacoustic facts have been considered (for details see Moore, 2014). Early results suggested that the level difference required for equal loudness of monaurally and diotically presented sounds was 10 dB. The subjective loudness of a sound doubles with about every 10 dB increase in physical intensity, and therefore it was assumed in (13) (12) 20 the early loudness model of Glasberg and Moore [31] that loudness sums across ears. However, later results suggested that the level difference required for equal loudness is rather between 5 to 6 dB. Glasberg and Moore therefore presented a new model to account for the lower dB values based on the concept of inhibition. Inhibition occurs when a strong input in one ear lowers or even stops, i.e. inhibits, the internal response evoked by a weaker input at the other ear (Moore [44]). Glasberg and Moore [31] implemented inhibition for binaural hearing by a gain function. Initially, the specific loudness pattern was smoothed with a Gaussian weighting function and the relative values of the smoothed function at the two ears were used to compute the gain functions of the ears. The gains were then applied to the specific loudness patterns at the two ears. The loudness for each ear was calculated by summing the specific loudness over the center frequencies and the binaural loudness was obtained by summing the loudness values across the two ears (Moore [44]). We used this procedure to calculate the binaural loudness values in this report. The binaural loudness model of Glasberg and Moore [31] has been implemented in PsySound3, a GUI-driven Matlab environment for analysis of audio recordings (http://www.psysound.org). Glasberg and Moore [31] assumed that the loudness of a brief sound is determined by the maximum of the short term loudness, while the long term loudness may correspond to the memory for the loudness of an event that can last for several seconds. For a time varying sound (e.g. an amplitude modulated tone) it is appropriate to consider the long time loudness as a function of time to calculate the time varying loudness. However, in this report, as the stimuli presented to the participants were noise bursts and can be considered steady and brief, we follow the assumption of Glasberg and Moore [31] of using the maximum of short time loudness as a measure of the loudness of the recordings. The means of the maxima values of Short Term Loudness in sones for the 10 versions for the 5, 50 and 500 ms recordings in the rooms of SN2010 [6] and SNG2016 [7] are presented in Tables 3, 4 and 5, respectively. From these tables, one sees that the loudness difference between the recordings without the object and with the object at 100 cm was less in the case of the lecture room of SNG2016 [7] than for the anechoic or conference room in SN2010 [6]. This may explain the low performance of the participants in the lecture room of SNG2016 [7]. The loudness values follow the same pattern as the sound pressure level analysis of the room acoustics chapter (Tables 2 and 5). However, the values in Tables 3 to 5 are psychophysical and depict not only the acoustics of the rooms but do also account for aspects of human hearing that are important for human echolocation. A comparison of the loudness results with the echolocation of persons will be made in the section "Threshold values, absolute and difference, for echolocation in static situations based on auditory model analysis", where we relate these psychophysical values to the echolocation performances. Table 3. Means of the maxima of Short Term Loudness, in sones, of the 10 versions for the recordings in the anechoic and conference room in SN2010 [6] and in the lecture room in SNG2016 [7] with a 5 ms duration signal. The blank cells indicate that no recordings were made at those distances. For the SN2010 there were two series of recordings with no reflecting object. Object distance (cm) No Object, recording 1 No Object, recording 2 50 Schenkman and Nilsson [6] Conference Anechoic room 13.4 13.3 20.7 room 19.3 19.4 26.7 Schenkman, Nilsson and Grbic [7] Lecture room 15.5 21 100 150 200 300 400 500 20.2 24.4 17.2 16.2 14.4 13.3 13.4 13.4 21.5 19.7 20 19.5 45 45.1 69.6 55.7 47.6 45.1 45.2 45.0 Table 4. Means of the maxima of Short Term Loudness, in sones, of the 10 versions for the recordings in the anechoic and conference room in SN2010 [6] with a 50 ms duration signal. There were two series of recordings with no reflecting object. Schenkman and Nilsson [6] Object distance (cm) Anechoic room Conference room No Object, recording 1 No Object, recording 2 40.1 40 63.7 52.3 50 100 150 200 40.3 300 40.3 400 40.2 500 40.1 Table 5. Means of the maxima of Short Term Loudness, in sones, of the 10 versions for the recordings in the anechoic and conference room in SN2010 [6] and in the lecture room in SNG2016 [7] with a 500 ms duration signal. The blank cells indicate that no recordings were made at those distances. For the SN2010 there were two series of recordings with no reflecting object. Schenkman and Nilsson [6] Anechoic Conference room 48.1 48.1 76.1 62.2 room 52.4 52.5 78.7 63.6 48.4 48.4 48.2 48.1 54.6 52.4 52.6 52.5 Schenkman, Nilsson and Grbic [7] Lecture room 52 54.7 52.5 Object distance (cm) No Object, recording 1 No Object, recording 2 50 100 150 200 300 400 500 22 Pitch analysis: autocorrelation with dual profiles Repetition pitch is a percept that underlies human echolocation for detecting objects. It is usually experienced as a coloration of the sound, perceived at a frequency equal to the inverse of the delay time between the sound and its reflection (Bilsen [14]; Bilsen and Ritsma [25]; see also Bassett and Eastmond [22]). As mentioned in the ACF section of Method, Room acoustics, in SN2010 and SNG2016 the reflecting objects were at distances of 50, 100, 150, 200, 300, 400 and 500 cm (although not all distances were used in both studies) resulting in delays of 2.9, 5.8, 8.7, 11.6, 17.4, 23.2 and 29 ms, where the repetition pitches would correspond to 344, 172, 114, 86, 57, 43 and 34 Hz, respectively. However, the actual delays might vary because of factors like the recording set up, speed of sound etc. and therefore the actual repetition pitch would be different. To test the presence of repetition pitch at these frequencies together with how this information would be represented in the auditory system, we used the PCP, BMM and NAP modules of the AIM, summarily presented above, to analyze the recordings from SN2010 [6] and SNG2016 [7]. The perception of repetition pitch can be created by presenting iterated rippled noise stimuli. The peaks in the autocorrelation function of these sounds are seen as the basis for repetition pith (Yost [26]; Patterson et al [45]). Hence, instead of the strobe finding and the temporal integration modules in AIM, we used the autocorr module as the final stage in our analysis to quantify repetition pitch information. Analysis by autocorrelation provides a feasible way to quantify repetition pitch, which we need to explain echolocation. We chose not to use the strobe temporal integration as the final stage, but it does not exclude that this might be how pitch information for echolocation is represented in the auditory system. To determine whether it is autocorrelation or strobe temporal integration that better explains repetition pitch perception and possibly also physiological processes involved in the auditory system, further experiments and analysis are needed. For the interested reader, we present as an example, some results obtained using the strobe temporal integration module for a 500 ms signal, see S.2 Supporting Information. After generating the ACF with the autocorr module in AIM, it has a dual profile development module, which sums up the ACF along both the temporal and the spectral domain. These features are relevant for human hearing in depicting how temporal and spectral information might be represented and are useful for analyzing repetition pitch. We therefore used this module to analyze the temporal and spectral results. The dual profile module plots in a single plot both the temporal and the spectral sum on frequency axis. The temporal profile and the spectral profile were scaled for this, and the inverse relation of time versus frequency (f = 1/t) was used to plot both time and frequency on a frequency scale. As an example, the dual plot for a 200 Hz pure tone is shown in Fig 9. All the recordings were not analyzed in this way. The recordings with the object at 300 to 500 cm in study SN2010 [6] and with 2, 4, 8, 16, 32 and 64, 5 ms clicks in study SNG2016 [7] do not provide any additional information for the module and were therefore not included in this autocorrelation analysis. 23 Fig 9. The autocorrelation function of the autcorr module in the AIM for a 200Hz pure tone, where the sum of the temporal axis (spectral profile, dashed line) and the frequency axis (temporal profile, unbroken line) can be seen. The peak in the temporal profile is at 200Hz. The peak in the spectral profile is approximately above 200Hz, this is due the high pass asymmetric function used in the dynamic compressive gammachirp filter, whose centre frequency shifts up as stimulus level increases. The temporal profile (unbroken line in the figures below) was calculated by summing the ACF output along 100 critical bands (50Hz to 8000Hz) at each time delay. The spectral profile (dashed line in the figures below) was calculated by summing the ACF output in each critical band along a 35 ms time delay. Therefore, the temporal profile consists of the sum of 100 critical bands at every sample of 35 ms time delay, and the spectral profile consists of the sum of 35 ms delay samples at every critical band. In the two studies that we analyze, the recordings had been presented to the participants with durations of 5, 50 or 500 ms plus an additional 450 ms of silence. Therefore, we had the same presentation duration, i.e. the whole signal was analyzed. For example, a 5 ms recording had 5 ms duration plus 450 ms of silence. However, when presenting the figures graphically, we use the first 70 ms time interval of the recordings. The dual profiles are presented for the recordings for the three signal durations and the two rooms in study SN2010 [6] and for the 5 ms and 500 ms signal in study SNG2016 [7]. It is important to note that the amplitude scale of the y-axis is different in each sub figure of each figure. The investigated attribute here is pitch and each sub figure with reflecting object should be compared with the sub figure with No object for each condition. A distinct peak in a sub figure which is absent in the sub figure with No object indicates the potential occurrence of the perception of a pitch. One should remember that the visual impression of a peak in a sub figure with No object may misleadingly indicate an auditory peak, unless one observes the different scales on the different y-axis for the different sub figures. The next section will deal with how to select peaks based on their peak strength. As mentioned above, the theoretical frequency of the repetition pitch for recordings with object at 100, 150 and 200 cm is 172, 114 and 86 Hz. The analysis of the 5 ms recordings show peaks approximately at these frequencies. For example, Fig 10c, 11c, 12b had peaks (marked by arrows in the subfigures) approximately at 172 Hz; Fig10d, 11d had peaks (marked by arrows in the subfigures) approximately at 82 Hz; Fig12c had a peak (marked by an arrow in the subfigure) approximately at 114 Hz. One reason that the peaks were not exactly at the theoretical values is probably due to the experimental setups of SN2016 and SNG2017 and of the room acoustics. Fig10b, 11b had no peaks at their 24 corresponding theoretical frequencies. However, this is due to a wider range in the y axis scale (cf Fig 10, Fig11 y axis labels). The spectral profiles on the other hand did not have peaks any closer to the theoretical frequencies (cf Fig10-12 dashed lines). There were small spectral differences but these may provide timbre information but not pitch information. In the 50 ms and 500 ms signal recordings, distinct peaks that could account for pitch perception were absent in the spectral profiles (Figs 13 to 17). We conclude that the spectral profiles (dashed line) did not provide information for pitch perception. The temporal profiles in Figs 13 to 17 might have some peaks approximately at the theoretical frequencies of the repetition pitch, but they were not clearly visible in these figures due to the scaling of the figures. Therefore, it is too premature to conclude that the temporal profile (unbroken line) is necessary for detecting the objects, if based on repetition pitch. A further analysis was therefore needed which quantified the peaks in the temporal profile. To determine the role of temporal information for detecting objects based on repetition pitch, the pitch strength development module of AIM was used. It measures the pitch perceived based on the peak strength. We elaborate this in the next section. The temporal profiles will be shown to have peaks at the theoretical frequencies of repetition pitch which, we believe, explains the perception of repetition pitch and thus also a major cause for detection by echolocation of the reflecting objects in the two studies. Fig 10. The dual profile of a 5 ms signal recorded in the anechoic room (in SN2010 [6]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B, C and D are for the recordings with object at 50, 100 and 200 cm respectively. 25 Fig 11. The dual profile of a 5 ms signal recorded in the conference room (in SN2010 [6]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B, C and D are for the recordings with object at 50, 100 and 200 cm respectively. Fig 12. The dual profile of a 5ms signal recorded in the lecture room (SNG2016 [7]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using 26 the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B and C are for the recordings with object at 100 and150 cm respectively. Fig 13. The dual profile of a 50 ms signal recorded in the anechoic room (SN2010 [6]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B, C and D are for the recordings with object at 50, 100 and 200 cm respectively. 27 Fig 14. The dual profile of a 50 ms signal recorded in the conference room (SN2010 [6]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B, C and D are for the recordings with object at 50, 100 and 200 cm respectively. Fig 15. The dual profile of a 500 ms signal recorded in the anechoic room (SN2010 [6]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B, C and D are for the recordings with object at 50, 100 and 200 cm respectively. 28 Fig 16. The dual profile of a 500 ms signal recorded in the conference room (SN2010 [6]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B, C and D are for the recordings with object at 50, 100 and 200 cm respectively. Fig 17. The dual profile of a 500 ms signal recorded in the lecture room (SNG2016 [6]). The solid line is the sum of the ACF along the spectral axis and the dashed line is the sum of the ACF along the time delay axis for 70 ms time interval. The temporal and spectral profiles 29 are scaled to be compared to each other. The 'x' axis is changed in the temporal profile by using the inverse relationship of time to frequency, f=1/t. The sub figure A is for the no object recording and sub figures B and C are for the recordings with object at 100 and 150 cm respectively. Pitch strength The peaks in the temporal profile of the autocorrelation function that we computed with the dual profile module of AIM were distributed without apparent order or meaning. It is not obvious which peak that corresponds to a pitch. There is a solution: The AIM model has a pitch strength module which calculates the pitch strength to determine if a particular peak is random or not. This module first calculates the local maxima and their corresponding local minima. The ratio of peak height to the peak width of the peak (local maxima) is subtracted from the mean of the peak height between two adjacent local minima to obtain the pitch strength (PS) of a particular peak. Two modifications were made by us in the pitch strength algorithm of AIM to improve its performance for the analysis. (1) The low pass filtering was removed as it smooths out the peaks and, (2) the pitch strength was measured with Equation (14). Figure 18 illustrates the pitch strength algorithm that was used. The peak with the greatest peak height has the greatest pitch strength and would be the perceived frequency of repetition pitch. PS = PH− PHLM�������� (14) where PS is the calculated pitch strength, PH is the height of the peak and PHLM is the mean of the peak height between two adjacent local minima. Fig 18. An illustration of the pitch strength measure computed using the pitch strength module of the AIM. The dot indicates the local maxima and the two stars are the corresponding local minima. The vertical dashed line is the pitch strength calculated using Equation (14). The frequency in Hz was computed by inverting the time delay, f = 1/t. The results of the calculated pitch strength for recordings of studies SN2010 [6] and SNG2016 [7] are presented in Tables 6 to 8. As can be seen, peaks were misleadingly identified for recordings without an object, which would not have caused a pitch perception. This happens because the pitch strength algorithm identifies all local maxima and minima in a 30 sound and thus also calculates the pitch strength for all random peaks (that have local maxima). Table 6. Mean of the pitch strength (autocorrelation index) of the 10 versions for recordings in SN2010 [6] and in SNG2016 [7] with the 5 ms duration signal. The blank cells indicate that no recordings were made at those distances. Schenkman and Nilsson [6] Object distance (cm) No Object, recording 1 No Object, recording 2 50 100 150 200 Anechoic room Conference room 0.37 0.40 2.54 1.06 - 0.78 0.80 9.65 2.43 - 0.35 0.85 Schenkman, Nilsson and Grbic [7] Lecture room 0.19 - - 0.29 0.28 - Table 7. Mean of the pitch strength (autocorrelation index) of the 10 versions for recordings in SN2010 [6] with the 50 ms duration signal. Schenkman and Nilsson [6] Object distance (cm) Anechoic room Conference room No Object, recording 1 No Object, recording 2 50 100 150 0.47 0.44 4.17 1.67 0.42 0.55 0.52 6.59 2.22 0.54 Table 8. Mean of the pitch strength (autocorrelation index) of the 10 versions for recordings in SN2010 [6] and in SNG2016 [7] with the 500 ms duration signal. The blank cells indicate that no recordings were made at those distances. Schenkman and Nilsson [6] Schenkman, Nilsson and Grbic [7] Object distance (cm) No Object, recording 1 No Object, recording 2 50 100 150 200 Anechoic room Conference room Lecture room 0.71 0.78 4.75 2.44 - 0.70 0.84 0.90 7.74 2.91 - 1.35 1.30 - - 1.36 1.42 - The unit for pitch strength in our analysis is the autocorrelation index, as it is computed on the autocorrelation function. The tabulated data show that for the 5 ms and 50 31 ms duration signals the pitch strength was greater than 1 for the object distances of 50 and 100 cm in the anechoic and conference rooms of SN2010 [6] (Tables 6 and 7). For the 500 ms duration signal, the strength was greater than 1 at distances of 50 and 100 cm in the anechoic room and at the distances of 50, 100 and 200 cm in the conference room. The lecture room in SNG2016 [7] also had a pitch strength greater than 1 at this condition, but the computed pitch strength was not consistent over a single frequency and it lasted for only 4 to 8 time frames. (The time frames had 35ms time delay computed from a 70 ms interval NAP signal. Each frame had a hop time of 10ms). This was not the case for the anechoic and conference rooms in SN2010 [6] that both had high pitch strengths at a particular frequency and lasted for 14 to 18 time frames of 35 ms intervals, each with a hop time of 10 ms. Additionally, in the lecture room in SNG2016 [7] with a reflecting object present, the pitch strength was not much different from when there was no object. This further illustrates the echolocating difficulties of the test persons in that study. The perceptual results of SN2010 [6] showed that the participants were able to detect the objects with a high percentage correct at object distances 50 and 100 cm in the anechoic room and at 50, 100 and 200 cm in the conference room (Schenkman and Nilsson [6]). As presented in the previous paragraph, the pitch strength was greater than 1 at these conditions. Pitch seems to be the most important information that listeners use to detect objects at these distances (see e.g. Schenkman and Nilsson [9]). Therefore, these results imply that there might be a perceptual threshold approximately equal to 1 (autocorrelation index) for pitch strength in echolocating situations. The peak with that pitch strength must exist for certain time frames for a person to perceive a repetition pitch. This is also dependent on the acoustics of the room. Relating the pitch strength results with the performance of the participants in the two studies SN2010 [6] and SN2016 [7] will be made in a following section on threshold values. Before this we shall analyze the results in terms of a timbre property, namely sharpness. Sharpness analysis In the Method part for analysis of room acoustics we described how the spectral centroid was used as a measure for timbre perception. The spectral centroid was computed on the time varying Fourier Transform. To address more specifically how human hearing perceives timbre, Fastl and Zwicker [33] computed the weighted centroid of the specific loudness rather than of the Fourier Transform. This psychophysical measure is called sharpness and indicates how sound extends from being perceived to vary from dull to sharp. We conducted the sharpness analysis for our recordings using code available from Psysound (Cabrera, Ferguson, and Schubert [35]). Sharpness varies over time and therefore the median was used as a representative measure for the perceived sharpness. The results of the mean of the medians of the perceived sharpness over the 10 versions in the anechoic, conference and lecture room in SN2010 [6] and SNG2016 [7] for 5, 50 and 500 ms duration signals are presented in Tables 11 to 14. The unit for sharpness is called acum. There are, to our knowledge, only a few studies on thresholds of sharpness. Pedrielli, Carletti, and Casazza [46] found that their participants had a just noticeable difference for sharpness of 0.04 acum. You and Jeon [47] found in a study on refrigerator noise that their participants had a just noticeable difference for sharpness of 0.08 acum. Assuming that 0.04 acum is a threshold value for sharpness, then the results in Tables 9 to 11 show that the difference in median sharpness in SN2010 [6] was greater than threshold for the object at 50 and 100 cm compared to the recordings without the object. In SNG2016 [7], the differences between the recordings with and without object were smaller than in SN2010 [6], although greater than 0.04 acum. It is possible that at shorter distances (say less than 200 cm) Object distance (cm) No Object, recording 1 No Object, recording 2 50 100 150 200 300 400 500 1.89 1.9 2.05 2.14 1.92 1.91 1.89 1.89 1.97 1.98 2.03 2.03 2 2.01 1.98 1.99 1.85 1.78 1.83 32 repetition pitch and loudness information might be more relevant for providing echolocation information than sharpness information. In study SN2010 [6] with reflecting object at distances 200, 300, 400 and 500 cm for 5 ms (anechoic and conference rooms), 50ms (anechoic and conference rooms) and 500ms signal (conference room) durations, the recordings had differences in mean of median sharpness of less than 0.04 acum when compared to the recordings without an object. However, in the anechoic room of SN2010 for the 500 ms signal duration [6], the recordings with object at 400 cm and 500 cm had differences in sharpness approximately greater than 0.04 acum when compared to the recordings without the object (Tables 9 to 11). This is information that blind persons might use to detect and identify a reflecting object at longer distances than, say at 200 cm. We discuss this issue further in the next section. Table 9. Mean of the 10 versions of mean of median of the sharpness (acum) for the recordings in anechoic, conference and the lecture room with the 5 ms duration signal. Schenkman and Nilsson [6] Schenkman, Nilsson and Grbic [7] Anechoic room Conference room Lecture room Table 10. Mean of the 10 versions of mean of median of the sharpness (acum) for the recordings in anechoic, conference and the lecture room with the 50 ms duration signal. Object distance (cm) No Object, recording 1 No Object, recording 2 50 100 150 200 300 400 500 Schenkman and Nilsson [6] Anechoic room Conference room 1.89 1.9 2.07 2.14 1.91 1.9 1.87 1.88 1.89 1.89 1.96 1.95 1.94 1.91 1.92 1.89 33 Table 11. Mean of the 10 versions of mean of median of the sharpness (acum) for the recordings in anechoic, conference and the lecture room with the 500 ms duration signal. Schenkman and Nilsson [6] Conference Anechoic room 1.86 1.88 2.12 2.12 room 1.94 1.94 2.1 2.04 Schenkman, Nilsson and Grbic [7] Lecture room 2.07 2.2 2.11 Object distance (cm) No Object, recording 1 No Object, recording 2 50 100 150 200 300 400 500 1.89 1.86 1.83 1.84 1.97 1.95 1.95 1.94 Threshold values, absolute and difference, for echolocation in static situations, based on auditory model analysis Background The stimuli in the studies [6, 7] were presented to the participants in a two-alternative-forced- choice (2AFC) manner. The participants had to compare two sounds and detect the sound with the echo. This perceptual decision is based on information in attributes of the stimuli. If a person can detect an attribute of an object with a 75 percentage of correct response, this limit is usually termed the absolute threshold. The judgments are commonly based on only one source of information. In the experiments in [6, 7] two sources of information were used. The sounds were presented in a 2AFC manner, where the participants compared the information of two sounds, with and without object, and made the decision. This is a difference threshold (Gescheider [48]), which is the threshold at which the participant can make a discrimination with about 75 percentage of correct response. As regards echolocation, we define the absolute threshold as the value of a perceptual attribute (e.g. loudness, pitch or sharpness) of a recording with a reflecting object at which a person has a 75 percentage correct response. The difference threshold is thus the difference of responses between the recordings with and without object, respectively, at which the person has 75 percentage of correct response. The experimental procedure used in the experiments was 2AFC, and therefore the difference threshold is the relevant measure of the echolocation ability of the persons. However, as the difference threshold is calculated from absolute thresholds both are for clarity shown here. The procedure for finding the difference thresholds and the corresponding results are presented below. Non-parametric versus parametric modeling of psychometric function for threshold analysis Psychometric functions relate perceptual results to physical parameters of a stimulus. Commonly the psychometric function is estimated by parametric fitting, i.e. it is assumed that the underlying relationship can be described by a specific parametric model. The parameters of such a model are then estimated by maximizing the likelihood. However, the most correct parametric model underlying the description of the psychometric function is unknown. 34 Therefore, estimating the psychometric function based on the assumptions of a parametric model may lead to incorrect interpretations (Zychaluk and Foster [49]). To address this problem, Zychaluk and Foster [49] implemented a non-parametric model to estimate the psychometric function that does not require assumptions about the empirical state of the underlying phenomenon. The psychometric function is thus modeled locally without assuming a "true" underlying function. Since the true relationship for the variables that determine human echolocation is unknown, we chose the method proposed by Zychaluk and Foster [49] in our analysis of the perceptual data. Next is a brief description of the non- parametric model for estimating the underlying psychometric function. Thereafter our analysis of the perceptual results is presented. A generalized linear model (GLM) is usually used when fitting a psychometric function. It consists of three components: a random component from the exponential family, a systematic component η and a monotonic differentiable link function g, that relates the two. Hence, a psychometric function, P(x), can be modeled by using Equation (15). The parameters of the GLM are estimated by maximizing the appropriate likelihood function (Zychaluk and Foster [49]). The efficiency of the GLM relies on how much the chosen link function, g , approximates the "true" underlying function. (15) 𝜂𝜂(𝑥𝑥)=𝑙𝑙[𝑆𝑆(𝑥𝑥)] However, as mentioned, the true function one can never know for certain. In non- parametric modelling, instead of fitting the link function g, the function η is fitted using a local linear method. For a given point, x, the value η(u) at any point u in a neighborhood of x is approximated by Equation (16) (Zychaluk and Foster [49]). 𝜂𝜂(𝑢𝑢) ≈ 𝜂𝜂(𝑢𝑢)−(𝑢𝑢−𝑥𝑥)𝜂𝜂′(𝑥𝑥) (16) where η'(x) is the first derivative of η. The actual estimate of the value of η(x) is obtained by fitting this approximation to the data over the prescribed neighborhood of x. Two features are important for this purpose, kernel K and the bandwidth h. A Gaussian kernel is preferred, as it has unbounded support and is best for widely spaced levels. An optimal bandwidth can be chosen using plugin, bootstrap or cross validation methods (Zychaluk and Foster [49]). As no method is guaranteed to always work, to find the optimal bandwidth for the analysis we chose a bootstrap method with 30 replications. When the bootstrap method failed to find an optimal bandwidth, cross validation was used to establish the optimal bandwidth. Distance thresholds for object detection The psychometric function was initially fitted to the mean proportion of correct responses with respect to the distance to the reflecting object. Figures 19, 20 and 21 show both the non- parametric modeling (local linear fit) and the parametric modeling of the perceptual results for the blind participants in study SN2010 [6]. The mean percentage correct is plotted as a function of distance for recordings with 5, 50 and 500 ms signals in the anechoic and conference rooms. The link function used for the parametric modeling was the Weibull function. Visual inspection shows that this link function was not appropriate, since the fit does not correspond well with the perceptual results. As mentioned, if one knows the underlying link function for the psychophysical data, then the parametric fit is a better fit than the local non-linear fit, but for the data we are analyzing one cannot assume a particular link function. However, the local linear fit correlates well with the perceptual results. This demonstrates 35 some of the advantages of using non-parametric modeling for the purposes of the present investigation. We used the means of the proportion of correct responses of the participants for the psychometric fitting. If we had used the individual responses, the individual thresholds would vary but the local linear fit would probably still, we believe, be well correlated with the perceptual results. Therefore, the results in the remaining part of the present section will be based on the psychometric function using local linear fit for the mean proportion correct answers. We used the implementation of the non-parametric model fitting in Matlab by Zychaluk and Foster [49]. The local linear fit needs at least three stimulus values to make a mathematically valid fit. As the recordings in the lecture room (in SNG2016 [7]) had only two stimulus values, at 100 and 150 cm, it was not possible to make a psychometric fit for these recordings. Fig 19. The parametric (Weibull fit) and non-parametric (Local linear fit) modeling of the mean proportion of correct responses for the 5 ms recordings of the blind participants in SN2010 [6] as a function of distance: (A) anechoic chamber. (B) conference room. Fig 20. The parametric (Weibull fit) and non-parametric (Local linear fit) modeling of the mean proportion of correct responses for the 50 ms recordings of the blind participants in SN2010 [6] as a function of distance: (A) anechoic chamber. (B) conference room. 36 Fig 21. The parametric (Weibull fit) and non-parametric (Local linear fit) modeling of the mean proportion of correct responses for the 500 ms recordings of the blind participants in SN2010 [6] as a function of distance: (A) anechoic chamber. (B) conference room. Distance perception of an object is not a perceptual attribute that was presented directly to the participants in studies SN2010 [6] and SNG2016 [7]. Therefore, the distance threshold obtained from the psychometric fit is a derived quantitative threshold. The distance threshold is the distance at which a person may detect an object with a probability of 75%. As the fitted psychometric function is discrete, it was not always possible for the fit to have an exact value of 0.75. Therefore, the threshold values at the proportion of correct responses within the range of 0.73 to 0.75 were initially calculated and the mean of the threshold values was determined as the actual threshold. The distance threshold in study SN2010 [6] for which the blind and the sighted could detect the object using echolocation, with a proportion of correct responses between 0.73 and 0.75, are shown in Table 12. The distances at which the blind participants could detect the reflecting object were farther away than for the sighted in both rooms and for all three sound signals. The threshold is positively related to the signal duration for both groups, i.e. the longer durations give a longer range of detection. We can also see that the blind persons could detect objects farther away in the conference room for all signals, but for the sighted this was only the case for the 500 ms signal. In the original study SN2010 [6, table 3], the calculations of at what distance a blind or a sighted person might detect a reflecting object were based on a parametric approach, yielding in general lower distance values for thresholds than our non- parametric approach. In other words, our reanalysis of the data show a higher sensitivity of both blind and sighted, than the values in [6]. Table 12. Distance thresholds (cm) for duration, room, and listener groups in study SN2010 [6]. The threshold values were calculated from the psychometric function of the blind and sighted participants' responses at the mean percentage of correct responses values of 0.73 to 0.75. Sound duration (ms) 5 Sighted 130 121 Blind 150 158 50 500 Blind 166 176 Sighted 160 147 Blind 172 247 Sighted 166 207 Room Anechoic Conference 37 Loudness thresholds, absolute and difference, for object detection Loudness is also a source of information for detecting reflecting objects by echolocation (Schenkman and Nilsson [9]; Kolarik et al, 2014 [2]). A common psycho-acoustical measure to express loudness is in the unit of sones (Yost [50]; Gescheider [48]). Therefore, we used values in sones for the local linear fitting to determine the absolute and difference threshold of loudness, where blind and sighted could detect a reflecting object. As elsewhere, the criterion was detection with a percentage correct between 73 and 75%. The mean loudness values, and the mean percentage of correct responses calculated from study SN2010 [6] were used as inputs for the psychometric fit. The resulting absolute threshold values of loudness for detecting the object are presented in Table 13. The data show that the thresholds for loudness for the blind participants were lower compared to those of the sighted, roughly 1 sone less in the anechoic chamber and 2 sones less in the conference room. Table 13. Absolute threshold values of loudness (sones) for duration, room, and listener groups in study SN2010 [6]. The threshold values were calculated from the psychometric function of the blind and sighted participants' responses at the mean percentage of correct responses value of 0.73 to 0.75. Room Anechoic Conference Sound duration (ms) 5 Sighted 17.5 24.1 Blind 16.8 22.6 50 500 Blind 43.7 49.4 Sighted 45.1 53.1 Blind 52.9 53.6 Sighted 53.2 55.3 The values in Table 13 may misleadingly lead a reader to infer that the shortest sounds had the lowest threshold. This is not the case. If we look at tables 3 and 5 the recording without object for the 5 ms signal had a loudness of approximately 13 sones and the 500 ms signal had approximately 45 sones. Considering these values, it is more appropriate to use the difference rather than the absolute threshold, since the detection based on loudness in SN2010 [6], as a consequence of the 2AFC method that had been used, was based on a relative judgment, a comparison, and not on an absolute judgment. The difference threshold values were calculated by subtracting the absolute threshold values in Table 13 with their corresponding loudness values for the recording without object, see Table 14. Table 14. Difference threshold values of loudness (sones) for duration, room, and listener groups in study SN2010 [6]. The threshold values were calculated from the psychometric function of the blind and sighted participants' responses at the mean percentage of correct responses value of 0.73 to 0.75. Room Anechoic Conference Sound duration (ms) 5 Sighted 4.1 4.7 Blind 3.4 3.2 50 Sighted 5 8 Blind 3.6 4.3 500 Blind 4.8 1.1 Sighted 5.1 2.8 The data in Table 14 give a different and more realistic picture than the data in Table 13. The blind detect objects at lower loudness values for all sounds, and both groups could detect with lower relative loudness levels (loudness difference between the recording with and without object) for 500ms duration signals in the conference room. The loudness model used to compute the mean loudness was the same for both test groups, and therefore we conclude 38 that the low thresholds of the blind persons are an effect of their perceptual ability. This conclusion is further discussed in the Discussion part of this report. Pitch thresholds, absolute and difference, for object detection The absolute and difference threshold values of pitch strength, as calculated by the autocorrelation index for which the blind and the sighted test persons in study SN2010 could detect the reflecting object, are presented in Tables 15 and 16, respectively. We will first discuss the results in terms of the absolute thresholds, but the more appropriate conclusions will be based on the difference thresholds. The absolute threshold varies for blind and sighted persons depending on signal durations and room conditions. The blind, for all conditions, had lower thresholds, the pitch strength increased with signal duration and the thresholds were lower in the anechoic room. It is possible that for shorter duration signals, the person may be inattentive and miss the signal and thus also the pitch information. The performance (percentage of correct response) of the participants with 5 and 50 ms signals may thus not only be based on pitch strength but also on cognitive factors such as attention. Schenkman and Nilsson [9] showed that when pitch and loudness information were presented together, at distances up to 200 cm to the object, the participants' performance was almost 100 percentage correct. The 500 ms recordings with the object at 50 and 100 cm in study SN2010 [6] had almost 100 percent correct response for both the blind and the sighted. Therefore, for the 500 ms signal condition it is likely that the likelihood to miss a signal and its pitch information because of non-attention is lower, and the perceptual results of the participants are likely to be based mostly on pitch information. There are two possible theoretical ways to regard how the hearing system treats the ACF values. We will here focus the analysis on the 500 ms signal, since as noted above, the 5 ms and 50 ms signals may have cognitive aspects that could bias the auditory model analysis. (1) Based on the above reasoning, and if we assume that the auditory system analyses the pitch information absolutely i.e. it does not compare the peak heights in the ACF between the recordings (when presented in a 2AFC manner), then the results indicate that the absolute threshold for detecting a pitch based on an autocorrelation process should be greater than 1.10 and 1.23 (as indicated by the autocorrelation index for the 500 ms signal) for the blind and the sighted, respectively, as shown in Table 15. (2) On the other hand, if we assume that the auditory system analyses the pitch information relatively i.e. it compares the peak heights in the ACF between the recordings (when presented in a 2AFC manner) then the results indicate that the difference threshold for detecting the pitch based on autocorrelation should be greater than 0.27 and 0.49 (autocorrelation index) for the blind and the sighted, respectively, as shown in Table 16. For all cases, the values show that the blind persons could detect echo reflections of objects having lower peak heights in the ACF, than the sighted could. Table 15. Absolute threshold values of pitch strength (autocorrelation index) for duration, room, and listener groups in study SN2010 [6]. The threshold values were calculated from the psychometric function of the blind and sighted participants' responses at the mean percentage of correct response values of 0.73 to 0.75. Sound duration (ms) 5 Sighted 0.88 2.21 Blind 0.77 1.54 50 500 Blind 0.8 1.07 Sighted 0.96 1.69 Blind 1.1 1.14 Sighted 1.23 1.41 Room Anechoic Conference 39 Table 16. Difference threshold values of pitch strength (autocorrelation index) for duration, room, and listener groups in study SN2010. The threshold values were calculated from the psychometric function of the blind and sighted participants' responses at the mean percentage of correct response values of 0.73 to 0.75. Sound duration (ms) Room Anechoic Conference Blind 0.39 0.75 5 Sighted 0.50 1.42 50 500 Blind 0.35 0.54 Sighted 0.51 1.16 Blind 0.36 0.27 Sighted 0.49 0.54 Sharpness thresholds, absolute and difference, for object detection As discussed previously, timbre indicates how we experience various qualities of sounds. We chose to study one aspect of timbre, sharpness, as a potential information source for object detection by echolocation. In analog to the previous psycho-acoustical parameters, we calculated the absolute and difference threshold values of sharpness for which the blind and the sighted test persons in study SN2010 [6] could detect a reflecting object using echolocation with a correct response value of 0.73 to 0.75. For quantitative values for sharpness, we used the psychophysical unit acum (e.g. Fastl and Zwicker [33]). Tables 17 and 18 show that for the blind and sighted participants both their absolute and difference thresholds, the sharpness values were about the same. However, unlike loudness and pitch strength the sharpness information need not be greater in value for the participants to detect the object. For sharpness, a listener must distinguish between timbres. This may include cognition, involving e.g. memory processes. When a participant in SN2010 [6] and SNG2016 [7] were presented with two stimuli in a 2AFC method, they distinguished the recording with the object from the recording without the object by identifying the one with the higher loudness level, stronger pitch strength or both. However, when a person uses sharpness for echolocation, e.g. in a 2AFC method, it is not necessary that the recording with the reflecting object has the higher sharpness value. The recording with the reflecting object might sound duller, i.e. having a lower value of sharpness, than the recording without the object. A person might use this information to detect or identify an object. Table 17. Absolute threshold values of the mean of the median sharpness (in the unit acum) for duration, room, and listener groups in study SN2010 [6]. The threshold values were calculated from the psychometric function of the blind and sighted participants' responses at the mean percentage of correct response values of 0.73 to 0.75. Room Anechoic Conference Blind 1.97 2.01 Sound duration(ms) 5 Sighted 1.98 2.03 50 500 Blind 1.96 1.94 Sighted 1.98 1.94 Blind 1.94 1.97 Sighted 1.96 1.97 Table 18. Difference threshold values of the mean of the median sharpness (in the unit acum) for duration, room, and listener groups in study SN2010 [6]. The threshold values were calculated from the psychometric function of the blind and sighted participants' responses at the mean percentage of correct response values of 0.73 to 0.75. Sound duration(ms) 5 50 500 40 Room Anechoic Conference Blind 0.07 0.03 Sighted 0.08 0.05 Blind 0.06 0.05 Sighted 0.08 0.05 Blind 0.07 0.03 Sighted 0.09 0.03 Table 11 showed that the sharpness values for the 500 ms duration recordings, with reflecting object at 400 and 500 cm in the anechoic room, had smaller sharpness values than the recording without an object. Interestingly, two blind participants (no. 2 and 6) performed better at these conditions than all the remaining participants, i.e. their proportion correct were approximately 0.7, even at 400 and 500 cm. We looked deeper into the performance of these two high-performing echolocators by making a local linear fit for the proportion correct of these two persons and the sharpness values of the 500ms recordings in the anechoic chamber that were calculated and are shown in Table 11. Cross validation was used to find the bandwidth of the local linear fit kernel. Figs 22 and 23 show the corresponding local linear fits. One can see clearly that when the proportion correct was approximately equal to 0.7, there were two absolute threshold values for sharpness, one higher and one lower. If we consider the mean of the sharpness of the two no object recordings at this condition (i.e. anechoic room and 500ms signal duration) it was about 1.87 (Table 11). Hence the difference threshold for the blind participant 2 would be 1.94 -1.87 = 0.07 acum and 1.83-1.87 = -0.04 acum. Similarly, the difference threshold for the blind participant 6 would be 1.97-1.87 = 0.10 acum and 1.83-1.87 = -0.04 acum. Perceptually, this means that the two high-performing blind participants could detect the object even when the recording with the object was duller than the recording without an object. A more detailed discussion on the importance of sharpness information for human echolocation is presented in the Discussion part below. Fig 22. The non-parametric (local linear fit) modeling of the proportion of correct response for the 500 ms recordings of the 2nd blind participant in SN2010 [6] as a function of the mean of median sharpness. As the function is discrete, the values of proportion correct responses approximately equal to 0.7 are considered. 41 Fig 23. The non-parametric (local linear fit) modeling of the proportion of correct response for the 500 ms recordings of the 6th blind participant in SN2010 [6] as a function of the mean of median sharpness. As the function is discrete, the values of proportion correct responses approximately equal to 0.7 are considered. Discussion We wanted to elucidate factors that determine echolocation and the differences of echolocation ability between blind and sighted persons. Neuroimaging and psychoacoustic methods give us insight into the high echolocating ability of blind persons, but these methods do not necessarily reveal the information in the acoustic stimulus that determines echolocation (at least when the information is not known) and also not fully how this information is represented and processed in the human auditory system. The implementation of auditory models for human echolocation was intended to establish the information that determines echolocation ability and its variability, and how this information is represented and processed in the human auditory system. Signal analysis was conducted on the physical signals and is presented in the Method part on Room acoustics. The aim was to find the physical information that is used for echolocation and to analyze the effects of room acoustics on human echolocation. We studied in particular the sound pressure levels, auto correlations and spectral centroids. The analyses were performed on the recordings of studies SN2010 [6] and SNG2016 [7]. The results demonstrate, as expected, that the acoustics of the room affect the sounds and thereby the physical attributes that are associated with it. However, the information represented in the auditory system is complex and it uses this physical information for further processing through the auditory neural system. To understand better what takes place in the auditory system of a person using echolocation, we used what we consider are the most relevant auditory models for human echolocation, that today are available in the literature. We thus studied how the corresponding perceptual attributes of sound pressure level, auto correlation and spectral centroid are encoded in the human auditory system. The results of the auditory models suggest that repetition pitch, loudness and sharpness all provide potential information for people to echolocate at distances below 200 cm. The results also indicate that at longer distances sharpness information may be used for human echolocation. A detailed discussion of how loudness, pitch and sharpness are essential for human echolocation and how they might be represented in the auditory system is 42 presented below in sections "Echolocation and loudness", "Echolocation and pitch" and "Echolocation and timbre". A discussion of how room acoustics and binaural information as well as movement affect human echolocation is presented in the section "Echolocation and room acoustics" and "Echolocation, binaural information and movement". We finally conclude by some comments on using auditory models for understanding human echolocation together with some theoretical implications. Echolocation and loudness Of the existing loudness models, we chose the one by Glasberg and Moore [31], since it has a good fit to the equal loudness contours in ISO 2006. The results of the model were related to the proportion of correct responses of the listeners in studies SN2010 [6] and SNG2016 [7] for calculating estimates of threshold values based on loudness. The loudness values are tabulated in Tables 3 to 5 and the resulting threshold values for detecting a reflecting object, when based on the percentage correct, are shown in Tables 13 and 14. The difference in loudness level between the loudness threshold and loudness level of the recordings without the object for 5, 50 and 500 ms duration signals in the anechoic room were approximately 4.2, 5 and 5 sones for the sighted persons. As an example, the 5 ms signal in the anechoic room Table 13 had a threshold of 17.5 sones for the sighted persons, while the mean for loudness with no object in this room was 13.3 sones, as shown in Table 3. The difference of these two values, 17.5 and 13.3, is 4.2 sones. For the conference room those differences were 5, 8 and 3 sones, respectively (Tables 3 to 5 and Table 13). These differences in loudness level make it possible for persons to echolocate, making loudness a potential information source for echolocation (see also Schenkman and Nilsson [9]; Kolarik et al, [2]). Comparing the loudness thresholds of sighted and blind persons, the thresholds of the blind persons were lower than those of the sighted test persons (Table 13). If loudness information is encoded in the same manner for both groups of test persons, which we believe is a reasonable assumption, then this analysis shows that blind persons may echolocate at lower loudness levels than sighted persons. Echolocation and pitch Repetition pitch is one of the important information sources that blind people use to detect a reflecting object at shorter distances (e.g. Bilsen [14]; [6]). We studied how this information is represented in the auditory system. For this purpose, a dual profile analysis was performed and is presented above in the section "Pitch analysis: autocorrelation with dual profiles". The results suggest that repetition pitch can be explained by the peaks in the temporal profile rather than by peaks in the spectral profile of the autocorrelation function. This is in agreement with a study by Yost [26], where the peaks in the temporal domain of the autocorrelation function form the basis for explaining the perception of repetition pitch. However, the dual profile analysis was not sufficient to determine the strength of the perceived pitch as the peaks were more random in the temporal profile of the autocorrelation function. A measure of pitch strength was therefore used that showed if the peaks were random or not and thereafter computed the pitch strength (Equation 14). The means of the resulting pitch strengths are shown in Tables 6 to 8 and the thresholds for detecting objects in Tables 15 and 16. Only the pitch strength values obtained from the 500 ms duration recordings in Table 8 were considered, as these recordings are not likely to be influenced by cognitive factors. The pitch strength threshold was lower for blind persons than for sighted. A reasonable assumption is that pitch information is encoded in the same manner for both blind and sighted persons. Then it appears that blind persons may echolocate with a lower pitch strength than sighted persons. The auditory models were used without changing its parameters 43 for the analysis of the two groups and it is thus not possible to infer what factors that determine the perceptual differences. Echolocation and timbre To determine to what extent that sharpness is useful for echolocation, we computed the weighted centroid of the sounds from studies SN2010 [6] and SNG2016 [7] for specific loudness using the code of Psysound3. Pedrielli, Carletti, and Casazza [46] showed in their analysis that the just noticeable difference for sharpness in their study was 0.04 acum. We used this value as a criterion of sharpness for detecting reflecting objects. The results from our analysis (Tables 9 to 11) show that the difference in sharpness was greater than 0.04 acum for recordings with the object at distances of 50, 100, 150 and 200 cm. However, at these distances both loudness and pitch information are more prominent. Hence, at distances shorter than 200 cm, sharpness might not be a major information source for echolocation. One may note that, in study SN2010 [6], for the 500 ms recording in the anechoic chamber, with the reflecting object at 400 cm and 500 cm, the sharpness difference was approximately equal to 0.04 acum when compared to the mean of the two recordings without the object (Table 11). A few of the blind test persons in SN2010 were able to detect objects at 400 cm. If the just noticeable difference for sharpness of 0.04 acum as found by Pedrielli, Carletti, and Casazza [46] also is a difference threshold for sharpness when echolocating sounds at longer distances than about 2 m, then sharpness can be used as vital information for blind people to detect objects at 400 cm. Here we point out that Tables 17 and 18 present a linear relationship between sharpness and percentage correct, i.e. if there is a higher value of sharpness then a higher probability of detection will result. However, as discussed previously, in distinction to loudness or pitch, sharpness does not need to be larger to indicate detection, since this may be indicated by either being perceived as dull or sharp. A psychometric function cannot depict this. An experiment controlling for sharpness information of the sound could clarify its role for echolocation. Echolocation and room acoustics Loudness, pitch and sharpness provide people with information useful for echolocation, but the efficiency of these also depend on the acoustics of the room and the character of the sounds. The results of many studies of human echolocation are evidence of this (Kolarik et al [2]; Tonelli, Brayda and Gori [52]; Vercillo et al [53]). The conference room in SN2010 [6] increased pitch strength and hence enabled the participants to echolocate at farther distances, while the lecture room in SNG2016 [7] decreased pitch strength and the listeners presumably had to rely more on other kinds of information like loudness, resulting in deterioration of object detection. The physical cause for the deterioration was probably because in SN2010 [6] the loudspeaker was on the chest of the artificial head, while in SNG2016 [7] it was 1 m behind the artificial head. Another physical cause for the deterioration might be that the reverberation time in the conference room in SN2010 [6] was 0.4s, but in the lecture room of SNG2016 [7] it was 0.6s.Too little reverberation does not seem to be beneficial for human echolocation (see also Tonelli, Brayda and Gori [52]), but too much cannot neither be useful. We hypothesize that there might be an optimal amount of reverberation for successful echolocation. Careful design of room acoustics should improve the possibilities for echolocation by blind persons. The effects of room acoustics for echolocation are also provided by the recordings in the anechoic room in study SN2010 [6]. The recordings with object at all distances in this room, including those at 400cm and 500cm, had no other reflections than from the reflecting object. The slight sharpness differences that resulted might be used by very skilled listeners to detect an object. 44 Echolocation, binaural information and movement Binaural information may provide additional information for echolocation. Both inter-aural level differences and time differences provide information for echolocation. Papadopoulos et al. [54] argued that information for obstacle discrimination were found for the frequency dependent inter aural level differences especially in the range from 5.5 to 6.5 kHz. Nilsson and Schenkman [55] found that the blind people in their study used interaural level differences more efficiently than the sighted. There is evidence that self-generated sounds (e.g. Kellogg [56]; Rice, [57]), as well as binaural information (e.g. Dunai, Peris-Fajarnés and Brusola [58]) is beneficial for echolocation. The recordings of the analyzed studies in this report, SN2010 [6] and SNG2016 [7], had the reflecting object directly in front of the recording microphones of the dummy head, and very little binaural information was therefore provided to the test persons. It was thus not considered as a source of information in this report, although we did calculate the SPL values at each ear. As can be seen in Tables 1 and 2, the SPL values at both ears were very similar. We may here add that the static nature of the recordings might have resulted in a poorer echolocation of the test persons. In a real situation movement gives additional information. Blind persons may move their heads and body, or the object might be moving. Additionally, they would also use their own sounds. It is reasonable to conclude that such sounds, together with motion, offer more information to blind persons than static sounds that are provided by an external source or by an experimenter [2-3]. The present article is based on two studies, both using stationary objects and stationary listeners. Already Wilson [4] showed theoretical reasons for the benefits for echolocation that could be achieved through motion. The same conclusions are obtained by the findings of Bassett and Eastmond [22] of how pitch changes depending on the distance to the reflecting object. More recent findings by Rosenblum et al [59] showed advantages for walking when echolocating. Furthermore, self-motion has been found to be beneficial for echolocation (Wallmeier and Wiegrebe [5]). In a study on blind children walking along a path, Ashmead, Hill and Talor [60] found that the children could avoid a box by utilizing non- visual information. These authors stated that congenitally blind children could utilize some of the auditory information and that they could coordinate this information with functionally important behavior, such as goal-directed locomotion. Arias et al [61] saw echolocation as a combination of action and perception, while Thaler and Goodale [3] in their review stressed that echolocation is an active process. Our present analysis should be extended to also include studies of motion of objects or persons, as well as of self-generated sounds. The same kinds of auditory models, possibly with some modification, could then be used to test processes or hypothesis pertaining to the additional benefits of motion and of self-generated sounds for echolocation. Comments on echolocation In psychoacoustic experiments a sound is usually presented to participants in a controlled manner and the perceptual or behavioral responses are measured. This makes it possible to identify cause-effect relationships between stimuli and response. However, although the stimuli are presented in a controlled manner in e.g. echolocation studies, the underlying cause for the echolocation is not obvious, e.g. whether it is biological, perceptual or psychological. For example, in the study SN2010 [6], the blind participants were able to perform better than the sighted, but the main cause for the high performance is not evident, except that the cause is related to blindness. the auditory model approach to human 45 De Volder et al [62] were maybe the first to describe different brain activities in blind and sighted persons in distance tasks by using positron emission tomography. Other advanced scanning techniques like functional magnetic resonance imaging of the brain may locate areas in the brain that are activated when persons use echolocation. These techniques can describe physiological processes relating to or underlying echolocation abilities (Thaler et al [63]; Thaler et al [64]), but such analyses do not fully reveal how the information used by blind persons for echolocation is processed and represented in the auditory system, how it cognitively is perceived and analyzed. Recently there has also been some serious criticism against MRI studies, that their results may be inflated by false-positive rates (Eklund, Nichols and Knutsson [65]). To address the issues of representation and processing, we implemented a number of auditory models; the binaural loudness model of Moore and Glasberg [32], the auditory image model, AIM, of Patterson, Allerhand, and Giguere [21] and the sharpness model of Fastl and Zwicker [33]. We chose the loudness model of Moore and Glasberg [32] since it agrees well with the equal loudness contours of ISO 2006 and also gives an accurate representation of binaural loudness (Moore [44]). We chose the auditory image model, AIM, since instead of using two different modules to depict frequency selectivity and compression; it uses a dynamic compressive gammachirp filterbank (dcGC) module to depict both the frequency selectivity and the compression performed by the basilar membrane. The AIM model of Bleeck, Ives, and Patterson [34], was thus used to analyze the information provided by repetition pitch. This model is also physiologically inspired. Finally, we used the loudness model of Glasberg and Moore [31] to analyze sharpness. The sharpness information was obtained from the weighted centroid of the specific loudness (Fastl and Zwicker [33]). A general review of auditory processing models is given by Dau [8]. The signal analysis performed on the physical stimuli showed how sound pressure level, autocorrelation and spectral centroid varied with the recordings. The results with AIM showed that the peaks in the temporal information was the likely source for echolocation at shorter distances, and this explanation is thus in line with the analysis by Bilsen [14] and Yost [26] in how the perception of repetition pitch is represented in people, i.e. that the information necessary for pitch perception is represented temporally in the auditory system. The analysis performed with the sharpness model showed that blind participants in our analyzed studies could have used sharpness to detect objects at longer distances and that both temporal and spectral information are required to encode this attribute. Our analysis has some similarities to that of Rowan et al [66] in utilizing models to analyze the perception of level information. Similar to their analysis we saw cross channel cues or spectral spread information as relevant for object detection, which we used in our quantification of sharpness. The analyses with the auditory models we have used do not fully explain how information necessary for the high echolocation ability of blind persons is represented in the auditory system. In order to investigate by auditory models, the neural and physiological causes to the different echolocation abilities of blind and sighted people, the parameters of the models could be varied to fit participants´ perceptual results. However, we have assumed that the high echolocation ability was due mainly to a perceptual ability common to both groups. Therefore, the thresholds for the blind and the sighted persons were obtained by comparing the results of the auditory models with the perceptual results of both test groups in SN2010 [6] and SNG2016 [7]. As mentioned, the blind participants were consistently better than sighted when echolocating, and they had lower thresholds of detection than the sighted for all parameters that were studied. The analysis with the auditory models confirmed that repetition pitch and loudness are important information sources for people when echolocating at shorter distances, which is in agreement with earlier results (e.g. Schenkman and Nilsson [6, 9]; Kolarik, Cirstea, Pardhan, and Moore [2]). Sharpness is a candidate for being an important source for 46 echolocation both at short and long distance. Psychoacoustic experiments could determine the usefulness of sharpness and other timbre qualities for echolocation. Highest ecological validity is probably reached by experiments with real objects in actual environments, but laboratory studies are a viable alternative. Today simulations of rooms and objects provide another option for how to study human echolocation, see e.g. Pelegrin-Garcia, Rychtáriková and Glorieux [66]. Conclusions We used auditory models to analyze how information for human echolocation in static situations is represented and processed in the auditory system, i.e. when no movement is involved. We focused on three perceptual attributes. Two of these, loudness and pitch, are known to be important for human echolocation. The third attribute, sharpness, an aspect of timbre, we considered important for echolocation and was therefore also studied. We used a number of auditory models: The binaural loudness model of Moore and Glasberg [32], the AIM model of Bleeck, Ives, and Patterson [34], the loudness model of Glasberg and Moore [31], and for sharpness information we used the weighted centroid of the specific loudness, as formulated by Fastl and Zwicker [33]. The main results of our analysis are the following. At shorter distances between person and reflecting object, repetition pitch, loudness and also sharpness provide information to detect objects by echolocation. At longer distances, sharpness information might be used for the same purpose. This tentative conclusion has to be justified experimentally by varying in particular the sharpness characteristics of echolocation sounds. Our analysis confirmed that repetition pitch can best be represented in the auditory system by the peaks in the temporal profile rather than by the spectral profile (see also Yost [26]). As the median sharpness information is computed by using the centroid of the specific loudness varying over time, it is represented by both the spectral and temporal information. For our analysis we have assumed that the auditory information for both blind and sighted persons is represented and processed in the same way. However, this assumption may not be true. The high echolocation ability of the blind may be the outcome of physiological changes in the neural system, as some studies have indicated (Thaler et al [63]; Thaler et al [64]). To investigate this in further detail, one should change the parameters of the auditory models and analyze the results together with data from neuroimaging and psychoacoustic experiments. If it is established that the underlying ability of blind persons is because of certain physiological conditions, then the parameters of the auditory models can be varied until the results from the auditory models agree with the psychoacoustic results. We used different auditory models to analyze loudness, pitch and sharpness attributes. It was assumed that the high echolocation ability of many blind persons is a result of a general perceptual ability, and we therefore computed perceptual thresholds for the blind and the sighted persons by using the same models for both groups. The analysis showed that the blind had lower thresholds than the sighted and could echolocate at both lower loudness and lower pitch strength levels. As noted above, the recordings in SN2010 [6] and SNG2016 [7], that form the basis for our analysis were recorded in static positions, i.e. there was no movement between the person and the object. In real life, a blind person would likely be moving, or the object would be so. In addition, the person is often using his/her own sounds, which is advantageous for echolocation (e.g. Kellogg, [56]; Rice, [57]). When movement is provided, and self-produced sounds are used, we believe that the thresholds for the blind persons would be even lower. These ideas are in alignment with the concept of surplus information in Schenkman and Nilsson [6] that more information makes perceptual tasks easier to perform, while lack of information makes perception ambiguous and difficult. This 47 concept follows from Gibsons´s [67] theory of ecological perception. In summary, we have shown the importance of pitch, loudness and timbre for human echolocation. These three characteristics have to be further studied, but especially the role of timbre attributes like sharpness, needs a deeper understanding. Neuroimaging, psychoacoustic experiments, and signal analysis including auditory models, may help us to understand how information necessary for the high ability of many blind persons or with visual impairments is represented and perceived. Acknowledgment The original data were collected together with Mats E. Nilsson. This work was partially supported by the Swedish Council for Working Life and Social Research, (grant number 2008-0600) 12006) https://www.promobilia.se/, both to BS. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. Author contributions Conceived and controlled the study: BS VG. Performed the simulations: BS VG. Analysed the data: BS VG. Wrote the paper: BS VG. Funding acquisition: BS. Contributed analysis tools: VG. Developed the theory: BS VG. Wrote the paper: BS VG. References https://forte.se/, Promobilia number (grant 1. Stoffregen TA, Pittenger JB (1995) Human echolocation as a basic form of and by perception and action. Ecological Psychology 7: 181- 216. 2. Kolarik AJ, Cirstea S, Pardhan S, Moore BC (2014) A summary of research investigating echolocation abilities of blind and sighted humans. Hearing Research 310: 60-68. 3. Thaler L, Goodale MA (2016) Echolocation in humans: an overview. WIREs Cogn Sci 2016, doi:10.1002/wcs.1408. 4. Wilson JP (1967) Psychoacoustics of obstacle detection using ambient or self- generated noise. In: Busnel RG, editor, Animal Sonar Systems. Biology and Bionics, vol 1, (pp 89 – 114) Jouy.en-Josas, France: Laboratoire de Physiologie Acoustique, INRA-CNRZ. 5. Wallmeier L, Wiegrebe L (2014) Self-motion facilitates echo-acoustic orientation in humans. Royal Society Open Science, 1:140185. 6. Schenkman BN, Nilsson ME (2010) Human echolocation: Blind and sighted persons' ability to detect sounds recorded in the presence of a reflecting object. Perception 39: 483-501. 7. Schenkman BN, Nilsson ME, Grbic N (2016) Human echolocation: Acoustic gaze trains and acoustic noise. Applied Acoustics 106: 77-86. for burst http://dx.doi.org/10.1016/j.apacoust.2015.12.008. 8. Dau T (2008) Auditory processing models. In: Havelock D, Kuwano S, Vorländer M, editors, Handbook of Signal Processing in Acoustics. New York: Springer. 9. Schenkman BN, Nilsson ME (2011) Human echolocation: Pitch versus loudness information. Perception 49: 840-852. 48 10. ASA (1973) American national psychoacoustical terminology, s3.20–1973. New York: American Standards Association. 11. Moore BC (2013) An Introduction to the Psychology of Hearing. San Diego: 12. ASA (1960) Acoustical terminology si, 1–1960. New York: American Standards Academic Press. Association. 13. De Cheveigné A (2010) "Pitch perception". In: Plack GJ, editor, Oxford Handbook of Auditory Science – Auditory Perception, Oxford University Press, Oxford, pp. 71–104. 14. Bilsen FA (1968) On the interaction of a sound with its repetitions. Doctoral dissertation, University of Technology, Delft, The Netherlands. 15. de Boer E (1956) On the "residue" in hearing. Academic thesis, Amsterdam. 16. de Boer E (1976) On the "residue" and auditory pitch perception. In Keidel WD, Neff WD, editors, Handbook of Sensory Physiology, vol V-3. Berlin: Springer, pp. 479-583. 17. Licklider JC (1951) A duplex theory of pitch perception. Cellular and Molecular Life Sciences 7: 128–134. 18. Goldstein JL (1973) An optimum processor theory for the central formation of the pitch of complex tones. The Journal of the Acoustical Society of America, 54:1496–1516. 19. Wightman FL (1973) The pattern-transformation model of pitch. The Journal of doi: 54: 407-416. America. of the http://dx.doi.org/10.1121/1.1913592. Acoustical Society 20. Terhardt E (1974) Pitch, consonance, and harmony. The Journal of the Acoustical Society of America 55: 1061-1069. 21. Patterson RD, Allerhand MH, Giguere C (1995) Time-domain modeling of peripheral auditory processing: A modular architecture and a software platform. The Journal of the Acoustical Society of America 98: 1890-1894 22. Bassett IG, Eastmond EJ (1964) Echolocation: Measurement of pitch versus distance for sounds reflected from a flat surface. The Journal of the Acoustical Society of America 36: 911-916. 23. Small AM, jr, McClellan ME (1963) Pitch associated with time delay between two pulse trains. The Journal of the Acoustical Society of America 35: 1246-1255. 24. Bilsen F (1966) Repetition pitch: monaural interaction of a sound with the repetition of the same, but phase shifted sound. Acustica 17: 295–300. 25. Bilsen F (1969) Repetition pitch and its implications for hearing theory. Acustica 26. Yost WA (1996) Pitch strength of iterated rippled noise. The Journal of the Acoustical Society of America 100: 3329-3335. 27. ANSI (1994) American national standard acoustical terminology, ansi s1.1-1994. New York: American National Standard Institute. 28. Plomp R (1970) Timbre as a multidimensional attribute of complex tones. In Plomp R, Smoorenburg GF, editors, Frequency Analysis and Periodicity Detection in Hearing, Leiden: W.J.Sitjthoff, pp. 397-414. 29. Plomp R (1976) Aspects of Tone Sensation: A psychophysical study. London, UK: Academic Press. 30. Plomp R, Steeneken, HJM (1969) Effect of phase on the timbre of complex tones. The Journal of the Acoustical Society of America, 46(2B), 409-421. 31. Glasberg BR, Moore BCJ (2002) A model of loudness applicable to time-varying sounds. Journal of Audio Engineering Society. 50.5: 331-342. 22: 63-73. 49 32. Moore BC, Glasberg BR (2007) Modeling binaural loudness. The Journal of the Acoustical Society of America. 121: 1604–1612 33. Fastl H, Zwicker E (2007) Psychoacoustics. Facts and Models. Berlin: Springer. 34. Bleeck S, Ives T, Patterson RD (2004) Aim-mat: the auditory image model in matlab. Acta Acustica United with Acustica 90: 781–787. 35. Cabrera D, Ferguson S, Schubert E (2007). 'Psysound3': Software for Acoustical and Psychoacoustical Analysis of Sound Recordings. Proceedings of the 13th International Conference on Auditory Display, Montréal, Canada, June 26-29. 36. Rowan D, Papadopoulos T, Edwards D, Holmes H, Hollingdale A, Evans L, Allen R (2013) Identification of the lateral position of a virtual object based on echoes by humans. Hearing Research 300: 56-65. 37. Peeters G, Giordano BL, Susini P, Misdariis N, McAdams S (2011) The timbre toolbox: Extracting audio descriptors from musical signals. The Journal of the Acoustical Society of America 130: 2902– 2916. 38. Patterson R D, Robinson K, Holdsworth J, McKeown D, Zhang C, Allerhand M (1992) Complex sounds and auditory images. Auditory Physiology and Perception 83: 429–446. 39. Irino T, Patterson RD (2006).A dynamic compressive gammachirp auditory Filterbank. IEEE Trans Audio Speech Lang Processing 14:2222-2232. 40. Patterson RD, Unoki M, Irino T (2003) Extending the domain of center frequencies for the compressive gammachirp auditory filter. The Journal of the Acoustical Society of America 114: 1529–1542. 41. Krumbholz K, Patterson RD, Nobbe A, Fastl H (2003) Microsecond temporal resolution in monaural hearing without spectral cues? The Journal of the Acoustical Society of America 113(5): 2790-2800. 42. Patterson, R. D. (1994). The sound of a sinusoid: Spectral models. The Journal of the Acoustical Society of America 96(3): 1409-1418. 43. Patterson, R. D. (1994). The sound of a sinusoid: Time-interval models. The Journal of the Acoustical Society of America 96(3): 1419-1428. 44. Moore BC (2014) Development and current status of the "Cambridge" loudness models. Trends in Hearing 18: 1–29. doi: 10.1177/2331216514550620. 45. Patterson RD, Handel S, Yost WA, Datta AJ (1996) The relative strength of the tone and noise components in iterated rippled noise. The Journal of the Acoustical Society of America 100: 3286-3294. 46. Pedrielli P, Carletti E, Casazza C (2008) Just noticeable differences of loudness and sharpness for earth moving machines. Proceedings of Acoustics08, June 29-July 4, 2008, Paris, France, pp. 1231-1236. 47. You J, Jeon JY (2008) Just noticeable differences in sound quality metrics for refrigerator noise. Noise Control Engineering Journal 56(6): 414-424. 48. Gescheider GA (1985) Psychophysics, 2nd ed. New Jersey: Lawrence Erlbaum. 49. Zychaluk K, Foster DH (2009) Model-free estimation of the psychometric function. Attention, Perception, and Psychophysics 71: 1414–1425. 50. Yost WA (2007) Fundamentals of hearing: An introduction. San Diego: Elsevier Academic Press. 51. Yost WA (1997) Pitch strength of iterated rippled noise when the pitch is ambiguous. Journal of the Acoustical Society of America 101: 644-1648. 52. Tonelli A, Brayda L, Gori M (2016) Depth echolocation learnt by novice sighted people. PLosOne 11(6):e0156654. Doi10.371/journal.pone.0156654. 50 53. Vercillo T, Jennifer L, Milne JL, Gori M, Goodale MA (2015) Enhanced auditory spatial localization in blind echolocators. Neuropsychologia 67: 35-40. 54. Papadopoulos T, Edwards DS, Rowan D, Allen R (2011) Identification of auditory cues utilized in human echolocation - objective measurement results. Biomedical Signal Processing and Control 6: 280-290 doi:10.1016/j.bspc.2011.03.005. 55. Nilsson ME, Schenkman BN (2016). Blind people are more sensitive than sighted people to binaural sound-location cues, particularly inter-aural level differences. Hearing Research 332: 223-232. 56. Kellogg WN (1962) Sonar system of the blind. Science 137: 399-404. 57. Rice CE (1969) Perceptual enhancement in the early blind. Psychological Research 19: 1-14. 58. Dunai L, Peris-Fajarnés G, Brusola F (2015) Virtual sound localization by blind people. Archives of Acoustics 40: 561-567. 59. Rosenblum LD., Gordon MS, Jarquin L. (2000) Echolocating distance by moving and stationary listeners. Ecological Psychology 12: 181-206. 60. Ashmead DH, Hill EW, Talor CR (1989) Obstacle perception by congenitally blind children. Perception & Psychophysics 46: 425-433. 61. Arias C, Bermejo F, Hüg MX, Venturelli N, Rabinovich D, Skarp AO (2012) Echolocation: an action-perception phenomenon. New Zealand Acoustics 25: 20– 27. 62. De Volder AG., Catalan-Ahumada M., Robert A, Bol A, Labar D, Coppens A, Michel C, Veraart C (1999) Changes in occipital cortex activity in early blind humans using a sensory substitution device. Brain Research 826: 128-134. 63. Thaler L, Arnott SR, Goodale, MA (2011) Neural correlates of natural human echolocation in early and late blind echolocation experts. PLoS One, 6, e20162. 64. Thaler L, Milne JL, Arnott SR, Kish D, Goodale MA (2014) Neural correlates of motion processing through echolocation, source hearing, and vision in blind echolocation and of Neurophysiology 111: 112-27 echolocation novices. Journal experts sighted 65. Eklund A, Nichols TE, Knutsson H (2016) Cluster failure: Why fMRI inferences for spatial extent have inflated false-positive rates. Proc Natl Acad Sci USA 113: 7900–7905. 66. Rowan D, Papadopoulos T, Archer L, Goodhew A, Cozens H, Guzman Lopez R, Edwards D, Holmes H, Allen R (2017) The detection of 'virtual' objects using echoes by humans: spectral cues. Hearing Research 350: 205 – 216. 67. Pelegrin-Garcia D, Rychtáriková M, Glorieux C (2017) Single simulated reflection audibility thresholds for oral sounds in untrained sighted people. Acta Acustica United with Acustica 103: 492 – 505. 68. Gibson JJ (1986). The Ecological Approach to Visual Perception. Hillsdale, NJ: Lawrence Erlbaum. 51 S1 Supporting information: Calculation of calibration constant The reference sound pressure level (SPL) for the calibration constants in the anechoic, conference and lecture rooms in SN2010 [6] and SNG2016 [7], were documented in dB(A), 77, 79 and 79 dB(A), respectively. The calibration constant of the recordings should, in principle, be A weighted. The results in Table S.1 show the calibrated levels calculated using the calibration constants in Equations A.1 and A.2, respectively. The A weighted signal gives an increase in calibrated levels by approximately less than 0.5 dB, which is negligible for human hearing. We therefore used Equation A.1, and not Equation A.2, for calculating the calibration constants for all recordings in this report. The results of Equations A.1 and A.2 for the calibrated levels with and without A weighting were calculated for the recordings of the 9th version of the left ear, 500 ms signal with no object of the first recording in the anechoic and conference rooms and of the 9th version of the left ear, 500 ms with no object recording in the lecture room, see Table S.1. Schenkman and Nilsson [6] Schenkman, Nilsson and Grbic [7] Anechoic room Conference room Lecture room Without A weighting: Equation A1 With A weighting: Equation A2 77.0 77.5 79.0 79.5 79.0 79.3 Table S.1: Calibrated levels with and without A weighting for the 9th version at the left ear for a 500ms sound with no reflecting object, first recording, in anechoic and conference rooms, and for the 9th version of left ear for 500ms sound with no reflecting object in lecture room. 20 𝐶𝐶𝐶𝐶= 10�𝑆𝑆𝑆𝑆𝑆𝑆−20∗𝑙𝑙𝑙𝑙𝑙𝑙10�𝑟𝑟𝑟𝑟𝑟𝑟(𝑟𝑟𝑠𝑠𝑙𝑙𝑠𝑠𝑠𝑠𝑙𝑙) 20∗10−6 � � 𝐶𝐶𝐶𝐶= 10�𝑆𝑆𝑆𝑆𝑆𝑆−20∗𝑙𝑙𝑙𝑙𝑙𝑙10�𝑟𝑟𝑟𝑟𝑟𝑟(𝐴𝐴𝐴𝐴𝐴𝐴𝑠𝑠𝑙𝑙ℎ𝑡𝑡𝑠𝑠𝑠𝑠𝑙𝑙(𝑟𝑟𝑠𝑠𝑙𝑙𝑠𝑠𝑠𝑠𝑙𝑙)) 20∗10−6 20 (A.1) � � (A.2) 52 S2 Supporting information: Pitch strength using strobe temporal integration The temporal profile of the stabilized auditory image for a recording of a 500 ms in the conference room in SN2010 [6] is shown in Fig S.2. As stated in the section "Auditory analysis of acoustic information", the stabilized auditory image was implemented with two modules, sf2003 and ti2003. A brief description of this implementation is given below. Fig S.2. The temporal profiles of the stabilized auditory image for a 500 ms signal recorded in the conference room (SN2010 [6]) at 495 ms time frame. The dot indicates the highest peak and the corresponding values indicate the pitch strength (calculated by Equation 14) and frequency in Hz (using the inverse relationship of time and frequency, f = 1/t).The sub figure A is for the no object recording and sub figures B, C and D are for the recordings with object at 50, 100 and 200 cm respectively. Initially, the sf2003 module uses an adaptive strobe threshold to issue a strobe on the NAP. After the strobe is initiated, the threshold initially rises along a parabolic path and then returns to the linear decay to avoid spurious strobes (cf. Fig 8). When the strobes have been computed for each frequency channel of the NAP, the ti2003 module uses the strobes to initiate a temporal integration. The time interval between the strobe and the NAP value determines the position where the NAP value is entered into the SAI. For example, if a strobe is identified in the 200Hz channel of the NAP at 5 ms time instant, then the level of the NAP sample at 5 ms time instant is added to the 1st position of the 200 Hz channel in the SAI. The next sample of the NAP is added to the 2nd position of the SAI. This process of adding the levels of the NAP samples continues for 35 ms and terminates if no further strobes are identified. In the case of strobes detected within the 35ms interval, each strobe initiates a temporal integration process. To preserve the shape of the SAI to that of the NAP, ti2003 uses weighting, viz. new strobes are initially weighted high (also the weights are normalized such that the sum of the weights is equal to 1) so that the older strobes contribute relatively less to the SAI. In this way the time axis of the NAP is converted into a time interval axis of the SAI. The temporal profile in the sub figures of Fig S.2 was generated by summing the SAI along the center frequencies. Fig S.2 shows that the recording with no reflecting object had a 53 pitch strength of 0.07, while the recording with the object at 200 cm (the fourth subfigure in Fig S.2) had a pitch strength of 0.1 at the corresponding frequencies of the repetition pitch. If this is the case for all the recordings has to be verified. Previous researchers (Yost [26]; Patterson et al [45]) analyzed the perception of repetition pitch by the autocorrelation function. We followed the same approach, since autocorrelation appears to be a good description of how repetition pitch is processed in the auditory system. To determine whether it is autocorrelation or strobe temporal integration that best accounts for human echolocation, repetition pitch and the relevant physiological processes of the auditory system, further analysis is needed, where these two concepts are studied and compared in a number of different conditions.
1908.00077
2
1908
2019-08-02T13:42:25
Learning in brain-computer interface control evidenced by joint decomposition of brain and behavior
[ "q-bio.NC" ]
Motor imagery-based brain-computer interfaces (BCIs) use an individuals ability to volitionally modulate localized brain activity as a therapy for motor dysfunction or to probe causal relations between brain activity and behavior. However, many individuals cannot learn to successfully modulate their brain activity, greatly limiting the efficacy of BCI for therapy and for basic scientific inquiry. Previous research suggests that coherent activity across diverse cognitive systems is a hallmark of individuals who can successfully learn to control the BCI. However, little is known about how these distributed networks interact through time to support learning. Here, we address this gap in knowledge by constructing and applying a multimodal network approach to decipher brain-behavior relations in motor imagery-based brain-computer interface learning using MEG. Specifically, we employ a minimally constrained matrix decomposition method (non-negative matrix factorization) to simultaneously identify regularized, covarying subgraphs of functional connectivity, to assess their similarity to task performance, and to detect their time-varying expression. Individuals also displayed marked variation in the spatial properties of subgraphs such as the connectivity between the frontal lobe and the rest of the brain, and in the temporal properties of subgraphs such as the stage of learning at which they reached maximum expression. From these observations, we posit a conceptual model in which certain subgraphs support learning by modulating brain activity in regions important for sustaining attention. To test this model, we use tools that stipulate regional dynamics on a networked system (network control theory), and find that good learners display a single subgraph whose temporal expression tracked performance and whose architecture supports easy modulation of brain regions important for attention.
q-bio.NC
q-bio
Learning in brain-computer interface control evidenced by joint decomposition of brain and behavior Jennifer Stiso1,2, Marie-Constance Corsi3,4, Jean M. Vettel5,2,6, Javier Garcia5,2, Fabio Pasqualetti7, Fabrizio De Vico Fallani3,4, Timothy H. Lucas9, and Danielle S. Bassett2,8,9,10,11,12,13 1Neuroscience Graduate Group, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104, USA 2Department of Bioengineering, School of Engineering & Applied Science, University of Pennsylvania, Philadelphia, PA 19104, USA 3Inria Paris, Aramis project-team, F-75013, Paris, France 4Institut du Cerveau et de la Moelle Epinire, ICM, Inserm, U 1127, CNRS, UMR 7225, Sorbonne Universit, F-75013, Paris, France 5Human Research & Engineering Directorate, US CCDC Army Research Laboratory, Aberdeen, MD, USA 6Department of Psychological & Brain Sciences, University of California, Santa Barbara, CA, USA 7Department of Mechanical Engineering, University of California, Riverside, CA 92521 8Department of Electrical & Systems Engineering, School of Engineering & Applied Science, University of Pennsylvania, Philadelphia, PA 19104, USA 9Department of Neurology, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104, USA 10Department of Psychiatry, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104, USA 11Department of Physics & Astronomy, College of Arts & Sciences, University of Pennsylvania, Philadelphia, PA 19104, USA 12The Santa Fe Institute, Santa Fe, NM 87501, USA and 13To whom correspondence should be addressed: [email protected] Motor imagery-based brain-computer interfaces (BCIs) use an individuals ability to volitionally modulate localized brain activity, often as a therapy for motor dysfunction or to probe causal rela- tions between brain activity and behavior. However, many individuals cannot learn to successfully modulate their brain activity, greatly limiting the efficacy of BCI for therapy and for basic scien- tific inquiry. Formal experiments designed to probe the nature of BCI learning have offered initial evidence that coherent activity across spatially distributed and functionally diverse cognitive sys- tems is a hallmark of individuals who can successfully learn to control the BCI. However, little is known about how these distributed networks interact through time to support learning. Here, we address this gap in knowledge by constructing and applying a multimodal network approach to decipher brain-behavior relations in motor imagery-based brain-computer interface learning using magnetoencephalography. Specifically, we employ a minimally constrained matrix decomposition method -- non-negative matrix factorization -- to simultaneously identify regularized, covarying sub- graphs of functional connectivity, to assess their similarity to task performance, and to detect their time-varying expression. We find that learning is marked by diffuse brain-behavior relations: good learners displayed many subgraphs whose temporal expression tracked performance. Individuals also displayed marked variation in the spatial properties of subgraphs such as the connectivity be- tween the frontal lobe and the rest of the brain, and in the temporal properties of subgraphs such as the stage of learning at which they reached maximum expression. From these observations, we posit a conceptual model in which certain subgraphs support learning by modulating brain activ- ity in regions important for sustaining attention. To test this model, we use tools that stipulate regional dynamics on a networked system (network control theory), and find that good learners display a single subgraph whose temporal expression tracked performance and whose architecture supports easy modulation of brain regions important for attention. The nature of our contribution to the neuroscience of BCI learning is therefore both computational and theoretical; we first use a minimally-constrained, individual specific method of identifying mesoscale structure in dynamic brain activity to show how global connectivity and interactions between distributed networks sup- ports BCI learning, and then we use a formal network model of control to lend theoretical support to the hypothesis that these identified subgraphs are well suited to modulate attention. 9 1 0 2 g u A 2 ] . C N o i b - q [ 2 v 7 7 0 0 0 . 8 0 9 1 : v i X r a INTRODUCTION Both human and non-human animals can learn to voli- tionally modulate diverse aspects of their neural activity from the spiking of single neurons to the coherent activity of brain regions [1 -- 3]. Such neural modulation is made possible by routing empirical measurements of the user's neural activity to a screen or other external display de- vice that they can directly observe [1, 4, 5]. Referred to as a brain-computer interface (BCI), this technology can be used to causally probe the nature of specific cognitive processes [6 -- 9], and offers great promise in the treatment of neural dysfunction [10 -- 13]. However, translating that promise into a reality has proven difficult [14 -- 16] due to the extensive training that is required and due to the fact that some individuals who undergo extensive training will only achieve moderate control [5, 17, 18]. A better under- standing of the neural processes supporting BCI learning is an important first step towards the development of BCI therapies and the identification of specific individu- als who are good candidates for treatment [17, 18]. While BCIs vary widely in their nature, we focus on the common motor imagery based BCIs where subjects are instructed to imagine a particular movement to mod- ulate activity in motor cortex. Performance on motor imagery based BCIs has been associated with a diverse array of neural features, demographic factors, and be- havioral measures [17 -- 21]. Neural features predicting performance are frequently identified in areas associated with either performing or imagining action; for example, better performance is associated with higher pre-task ac- tivity in supplementary motor areas [22] and larger grey matter volume in somatomotor regions [22]. Interest- ingly, performance has also been predicted by activity in a diverse range of other cognitive systems relevant for sustained attention, perhaps due to the high cognitive demands associated with BCI learning [18]. Specifically, better performance is associated with greater parietal power suppression in the α band, midline power suppres- sion in the β band, and frontal and occipital activation with motor power suppression in the γ band [20, 23, 24]. The role of sustained attention in BCI control is corrob- orated by the fact that personality and self-report mea- sures of attention predict successful learning [19]. The heterogeneity of predictors suggests the possibility that individual differences in the interactions between cogni- tive systems necessary for action, action planning, and attention might explain the idiosyncratic nature of BCI control, although these interactions are challenging to quantify [6, 25]. Assessing the interactions between cognitive systems has historically been rather daunting, in part due to the lack of a common mathematical language in which to frame relevant hypotheses and formalize appropriate computational approaches. With the recent emergence and development of network science [26], and its appli- cation to neural systems [27], many efforts have begun to link features of brain networks to BCI learning specif- ically and to other types of learning more generally. In this formal modeling approach [28], network nodes rep- resent brain regions or sensors and network edges repre- sent statistical relations or so-called functional connec- tions between regional time series [29]. Recent studies have demonstrated that patterns of functional connec- tions can provide clearer explanations of the learning process than activation alone [30], and changes in those functional connections can track changes in behavior [31]. During BCI tasks, functional connectivity reportedly in- creases within supplementary and primary motor areas [15] and decreases between motor and higher-order as- 2 sociation areas as performance becomes more automatic [32]. Data-driven methods to detect putative cognitive systems as modules in functional brain networks [33, 34] have been used to demonstrate that a particularly clear neural marker of learning is reconfiguration of the net- work's functional modules [31]. Better performance is accompanied by flexible switching of brain regions be- tween distinct modules as task demands change [35 -- 39]. While powerful, such methods for cognitive system de- tection are built upon an assumption that limits their conceptual relevance for the study of BCI learning. Specifically, they enforce the constraint that a brain re- gion may only affiliate with one module at a time [40], in spite of the fact that many regions, comprised of het- erogeneous neural populations, might participate in mul- tiple neural processes. To address this limitation, recent efforts have begun to employ so-called soft-partitioning methods that detect coherent patterns in mesoscale neu- ral activity and connectivity [40 -- 44]. Common examples of such methods are independent component analysis and principal component analysis, which impose pragmatic but not biological constraints on the orthogonality or in- dependence of partitions. An appealing alternative is non-negative matrix factorization (NMF), which achieves a soft partition by decomposing the data into the small set of sparse, overlapping, time-varying subgraphs that can best reconstruct the original data with no require- ment of orthogonality or independence [45]. Previous applications of this method to neuroimaging data have demonstrated that the detected subgraphs can provide a description of time varying mesoscale activity that com- plements descriptions provided by more traditional ap- proaches [40]. For example, some subgraphs identified with NMF during the resting state have similar spatial distributions to those found with typical module detec- tion methods, while others span between modules [40]. As a minimally constrained method for obtaining a soft partition of neural activity, NMF is a promising candi- date for revealing the time-varying neural networks that support BCI learning. Here, we investigate the properties of dynamic func- tional connectivity supporting BCI learning. In indi- viduals trained to control a BCI, we calculate single trial phase-based connectivity in magnetoencephalogra- phy (MEG) data in three frequency bands with stereo- typed behavior during motor imagery: α (7-14 Hz), β (15-25 Hz), and γ (30-45 Hz). We construct multimodal brain-behavior time series of dynamic functional connec- tivity and performance, and apply NMF to those time series to obtain a soft partition into additive subgraphs [45] (Fig. 1). We determine the degree to which a sub- graph tracks performance by defining the performance loading as the similarity between each subgraph's tem- poral expression and the time course of task accuracy. We first identify subgraphs whose performance loading predicted the rate of learning and then we explore the spatial and temporal properties of subgraphs to iden- tify common feature across participants. We hypothesize that subgraphs predicting learning do so by being struc- tured and situated in such a way as to easily modulate patterns of activity that support sustained attention, an important component of successful BCI control [18]. Af- ter demonstrating the suitability of this approach for our data (Fig. S1A-B), we test this hypothesis by capitaliz- ing on recently developed tools in network control theory, which allowed us to operationalize the networks ability to activate regions involved in sustained attention as the energy required for network control [46]. Collectively, our efforts provide a careful network-level description of neural correlates of BCI performance and learning rate, and a formal network control model that explains those descriptions. METHODS Participants Written informed consent was obtained from twenty healthy, right-handed subjects (aged 27.45 ± 4.01 years; 12 male), who participated in the study conducted in Paris, France. Subjects were enrolled in a longitudinal electroencephalography (EEG) based BCI training with simultaneous MEG recording over four sessions, spanning 2 weeks. All subjects were BCI-naive and none presented with medical or psychological disorders. The study was approved by the ethical committee CPP-IDF-VI of Paris. BCI task The BCI task consisted of a standard 1 dimensional, two-target box task (14) in which the subjects modulated their α [8-12 Hz] and/or β [14-29 Hz] activity to control the vertical position of a cursor moving with constant ve- locity from the left side of the screen to the right side of the screen. Both cursor and target were presented using the software BCI 2000 [47]. To hit the target-up, the subjects performed a sustained motor imagery of their right-hand grasping and to hit the target-down they re- mained at rest. Some subjects reported that they imag- ined grasping objects while others reported that they simply imagined clenching their hand to make a fist. Each trial lasted 7 s and consisted of a 1 s inter-stimulus interval, followed by 2 s of target presentation, 3 s of feedback, and 1 s of result presentation. In the present study, we restricted our analysis to the feedback portion of the motor imagery task because we were interested in the neural dynamics associated with learning to volition- ally regulate brain activity rather than in the neural dy- namics occurring at rest. BCI control features including EEG electrode and frequency were selected in a calibra- tion phase at the beginning of each session, by instructing the subjects to perform the BCI tasks without any visual feedback. The subjects were seated in front of a screen at a distance of 90 cm. To ensure a stable position of the 3 hands, each subject rested their arms on a comfortable support, with palms facing upward. We also recorded electromyogram (EMG) signals from the left and right arm of subjects, electrooculograms, and electrocardio- grams. EMG activity was manually inspected to ensure that subjects were not moving their forearms during the recording sessions. MEG Data Preprocessing As a preliminary step, temporal Signal Space Sepa- ration (tSSS) was performed using MaxFilter (Elekta Neuromag) to remove environmental noise from MEG activity. All signals were downsampled to 250 Hz and segmented into trials. ICA was used to remove blink and heartbeat artifacts. An FFT of the data from each subject was inspected for line noise, although none was found in the frequency bands studied here. We note that the frequency of the line noise (50 Hz) was out- side of our frequency bands of interest. In the present study, we restricted our analyses to gradiometer sen- sors. Gradiometers sample from a smaller area than magnetometers, which is important for ensuring a sep- arability of nodes as expected by network models [48]. Furthermore, gradiometers are typically less susceptible to noise than magnetometers [49]. We combined data from planar gradiometers in the voltage domain using the 'sum' method from Fieldtrip's ft combine planar() function (http://www.fieldtriptoolbox.org/). Connectivity Analysis To estimate phase-based connectivity, we calculated the weighted phase-locking index (wPLI) [50]. The wPLI is an estimate of the extent to which one signal consis- tently leads or lags another, weighted by the imaginary component of the cross-spectrum of the two signals. Us- ing phase leads or lags allows us to take zero phase lag signals induced by volume conduction and to reduce their contribution to the connectivity estimate, thereby ensur- ing that estimates of coupling are not artificially inflated [50]. By weighting the metric by the imaginary com- ponent of the cross spectrum, we enhance robustness to noise [50]. Formally, the wPLI between two time series x and y is given by E{imag(Γxy)} E{imag(Γxy)} , φ(x, y) = (1) where E{} denotes the expected value across estimates (here, centered at different samples), Γxy denotes the cross spectrum between signals x and y, and imag() se- lects the imaginary component. We first segment MEG data from gradiometers into 3 second trials, sampled at 250 Hz. The cross spectrum 4 FIG. 1. Schematic of non-negative matrix factorization. (1) MEG data recorded from 102 gradiometers is segmented into windows (w1, w2, w3, w4, ... wn) that each correspond to a single BCI trial. (2) In each window, functional connectivity is estimated as the weighted phase-locking index between sensor time series. The subject's performance on each trial is also recorded. (3) The lower diagonal of each trial (highlighted in grey in panel (2)) is reshaped into a vector, and vectors from all trials are concatenated to form a single data matrix. The subject's time-varying performance forms an additional row in this configuration matrix. (4) The NMF algorithm decomposes the configuration matrix (composed of neural and behavioral data) into m subgraphs (where m is a free parameter), with two types of information: (i) the weight of each edge in each subgraph, also referred to as the connection loading (viridis color scale), and (ii) the time varying expression of each subgraph (black line graphs). From these data, we calculate the performance loading, or how similar the time-varying performance is to each subgraph's expression (shades of purple). is then estimated using wavelet coherence [51] in each of three frequency bands of interest (α 7-14 Hz, β 15- 20 Hz, and γ 31-45 Hz), with wavelets centered on each timepoint. We chose to compute the wavelet coherence due to the fact that -- unlike Welch's method -- it does not assume stationarity of the signal [51]. We implemented the procedure in the Fieldtrip package in MATLAB, with a packet width of 6 cycles and zero-padding up to the next power of two ('nextpow2'). We then calculate the wPLI as the mean of the imaginary component of the cross spectrum, divided by the imaginary component of the mean of the cross spectrum. We then construct a network model of these statistical relationships where sensors (N = 102) are nodes, and the weight of the edge between node i and node j is given by the weighted phase-locking value. The graph, G, com- posed of these nodes and edges is a weighted, undirected graph that is encoded in an adjacency matrix A. By con- structing this network model, we can use statistics from graph theory and computational approaches from control theory to quantify the structure of inter-sensor functional relations [6, 28]. Uniformly Phase Randomized Null Model In order to ensure that our results are not due to choices in preprocessing, the time invariant cross- correlation of neural signals, or the autocorrelation of neural signals, we repeated all of the preprocessing and analysis steps with a uniformly phase randomized null model [52]. To enhance the simplicity and brevity of the exposition, we will also sometimes refer to this construct simply as the null model. Surrogate data time series from the null model were calculated using a custom function in MATLAB. Essentially, the FFT of the raw data is taken, the same random phase offset is added to every channel, and then the inverse FFT is taken to return the signal to the time domain [53]. Mathematically, this process is achieved by taking the discrete Fourier transform of a time series yv: V −1(cid:88) v=0 Y (u) = yvei2πuv/V , (2) where V is the length of the time series, v indexes time, and u indexes frequencies. We then multiply the Fourier transform by phases chosen uniformly at random before transforming back to the time domain: V −1(cid:88) v=0 yv = 1√ V eiau Y (u) e−i2πkv/V , (3) where the phase at ∈ [0, 2π). Construction of a Multimodal Configuration Matrix In this work, we wished to use a data-driven matrix decomposition technique to identify time-varying sub- graphs of functional connectivity that support learning. Specifically, we created a multimodal configuration ma- trix of edge weights and BCI performance over time, prior to submitting this matrix to a decomposition algorithm that we describe in more detail below. To construct the matrix, we first vectorize the upper triangle (not includ- ing the diagonal) of each trial's connectivity matrix, and then we concatenate all of the vectors and our one per- formance measure into an E × τ matrix, where τ is the number of trials (384, if no trials were removed), and E is the number of edges (5151) plus the number of behav- ioral measures (1). This concatenation process results in a 5152 × 384 multimodal (brain-behavior) matrix. In this task, each subject's performance is recorded as their 5 percentage of successful trials (out of 16) on each run, and we therefore interpolate the performance time series to obtain a graded estimate of percentage correct in each trial that is τ time points long. The performance vector is then normalized to have the same mean as the other rows of the configuration matrix. Non-negative Matrix Factorization We used a data-driven matrix decomposition method -- non-negative matrix factorization (NMF) -- to iden- tify time-varying groups of neural interactions and be- havior during BCI learning [45]. Intuitively, NMF de- composes a matrix into a set of additive subgraphs with time-varying expression such that a linear combination of these subgraphs weighted by temporal expression will recreate the original matrix with minimal reconstruction error [40, 45]. The NMF algorithm can also be thought of as a basis decomposition of the original matrix, where the subgraphs are basis sets and the temporal coefficients are basis weights. Unlike other graph clustering methods [54, 55], NMF creates a soft partition of the original net- work, allowing single edges to be a part of multiple sub- graphs. Additionally, unlike other basis decomposition methods [54, 55], NMF does not impose harsh constraints of orthogonality, or independence of the subgraphs; it simply finds the most accurate partition, given that the original matrix is non-negative. In many systems (in- cluding phase-locking) the non-negativity constraint is not difficult to satisfy, and is beneficial in physical sys- tems where the presence of a negative weight would be difficult to interpret. Formally, the NMF algorithm will approximate an E × T configuration matrix A by the multiplication of two matrices: W, the subgraph matrix with dimensions E × m, and H, with dimensions m × T . Here, E is the number of time varying processes (behavior and func- tional connections derived from MEG data), T is the number of time points, and m is the number of subgraphs. We solve for W and H such that: min W,H 1 2 A − WH2 F + αW2 F + β H(:, t)2 1 (4) T(cid:88) t=1 where β is the penalty to impose sparse basis weights, and α is the regularization for the basis set. Regularization is frequently used in machine learning algorithms to avoid overfitting data, which is especially important when em- ploying these techniques to examine highly variable single trial estimates of functional connectivity [56]. Addition- ally, selecting for sparsity will encourage the characteri- zation of local neural processes where many edges do not contribute [40]. From many such local processes arises the diversity of cognitive functions involved in complex tasks such as BCI control [57]. To solve the NMF equation, we use an alternating non- negative least squares with block-pivoting method with 100 iterations for fast and efficient factorization of large matrices, where W and H with non-negative weights are drawn from a uniform random distribution on the inter- val [0, 1][58]. The parameter m is drawn from the range (2,20), and α and β are drawn from the range (0.001,2). We select for parameters that will both minimize the residual error, and maximize the temporal and subgraph sparsity [40]. Specifically, we select the optimal parame- ters ¯m, ¯α, and ¯β that are in the lowest 25th percentile for residual error, and the highest 25th percentile for tempo- ral and subgraph sparsity. This procedure resulted in an average ¯m of 7.4, an average ¯α of 0.46, and an average ¯β of 0.45. Distributions of parameters and reliability across runs are shown in Fig. S2 and S3. Given the non-deterministic nature of this approach, we also test for the stability of our identified clusters us- ing a consensus clustering algorithm [59]. Our procedure was comprised of the following ordered steps: (1) run the NMF algorithm r = 100 times per multimodal configu- ration matrix, (2) concatenate the subgraph matrix W across r runs into an aggregate matrix with dimensions E × (r ¯m), and (3) apply NMF to the aggregate matrix to determine a final set of subgraphs and expression co- efficients [40]. While the implementation is heuristic in nature, we found that across two runs of the algorithm, we obtain highly consistent selections for parameters (see Supplement), bolstering confidence in the robustness of the subsequent analyses. Subgraph Inclusion Most subgraphs are sparse, with distributions of tem- poral coefficients skewed towards zero (see Fig. S4). However, for every subject and every frequency band, one subgraph showed very little regularization (no edges were equal to 0) and had a uniform, rather than skewed distribution of temporal coefficients. These subgraphs are clear outliers from the others, and appear to be cap- turing global phase-locking across the entire brain, rather than any unique subsystem. To answers questions about the time varying interactions between neural systems, we were interested in differences between the subgraphs that were spatially localized, having edges regularized to zero. Because including these outlier subgraphs would obscure those differences, we removed these subgraphs from all further analyses. Group Average Subgraphs After applying NMF to the multimodal brain-behavior matrix, we next turned to a study of the nature of the detected subgraphs after ranking them by performance loading. Specifically, we were initially interested in de- termining which edges contributed to each ranked sub- graph most consistently across the population. For this purpose, we used a consistency based approach to cre- 6 FIG. 2. BCI performance. Each subject's average perfor- mance across four days within two weeks. BCI Score is the percentage of correct trials during that session. ate a group representative subgraph for each ranked sub- graph [60]. In this procedure, each subject's subgraph was first thresholded to retain only the 25% strongest connections (see Fig. S5 for evidence that results are ro- bust to variations in this choice). We then constructed an average N × N subgraph G, where N is the number of channels and where each element Gij quantifies how many subjects (out of 20) displayed an edge between re- gion i and region j in their thresholded subgraph. In addition to visually depicting these group representative subgraphs, we also wished to summarize their content in large anatomical regions. We therefore binned edges into 10 anatomically defined areas in frontal, motor, parietal, occipital, and temporal lobes in both hemispheres. Lobes were obtained from BrainStorm [61] software (neuroim- age.usc.edu/brainstorm/Tutorials/MontageEditor) (Fig. S8). Optimal Control Our final broad goal was to provide a theoretical ex- planation for why certain networks support BCI learning. We hypothesized that these regularized networks might have structures that make it easier for the brain to mod- ulate the patterns of activity that are necessary for BCI control. This hypothesis motivated us to formulate and validate a model to explain how the sparse statistical re- lationships characteristic of each subgraph could support the production of brain activity patterns implicated in BCI learning [62, 63]. Additionally, this model should account for the brain's ability to reach these patterns of activity in the context of the BCI task, where there is increased volitional modulation of the left motor cortex. Here, we use tools from network control theory to sat- isfy these conditions [64]. Specifically, we characterize the theoretical brain activity at each sensor as a vector x(t), and we use the adjacency matrix A of a subgraph 40507090Session 1Session 2Session 3Session 4TimeBCI Score(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)(cid:31)405060708090Sess1Sess2Sess3Sess4TimeBCI Scoreas.factor(sess_cont)(cid:31)(cid:31)(cid:31)(cid:31)6080 to quantify the ease with which that activity can affect other regions. We then incorporate volitional input con- trol as input into the brain (u(t)) at a specific region (given by B). Then, by stipulating x(t) = Ax(t) + Bu(t), (5) we model the linear spread of activity along the connec- tions in A in the context of input to regions given in B. With this model of network dynamics, optimal control trajectories can be formalized and identified by develop- ing a cost function that seeks to minimize two terms: (i) the distance of the current state from the target state and (ii) the energy required for control. Specifically, we solve the following minimization problem: (xT − x(t))T (xT − x(t)) + ρuκ(t)T uκdt, (cid:90) T min u 0 7 node "b", an assumption that is violated by measures of statistical similarity such as the Pearson correlation coefficient which is the measure of functional connectiv- ity most commonly employed in neuroimaging studies. Because we are using neither structural connectivity nor common measures of functional connectivity, it was nec- essary for us to first prove that the networks we are study- ing are consistent with our model. To address the first point regarding the propagation of activity along edges, we demonstrate that the structure of the subgraphs used have utility in predicting empirical brain state transi- tions, and that the relative contribution of each subgraph is related to its temporal expression (Fig. S1C-D). It is only in light of these validations that we are able to in- terpret our results as a potential model for driving brain activity. To address the second point regarding isolation of pairwise relations not due to third party effects, we note that the matrix A that we study reflects statistical similarity in phase after strict regularization that removes redundant statistical relationships (Fig. S1A-B). s.t. x = Ax(t) + Bu(t), x(0) = x0, and x(T ) = xT , (6) Target state definition where ρ is a free parameter that weights the input con- straint, xT is the target state, and T is the control hori- zon, which is a free parameter that defines the finite amount of time given to reach the target state. Dur- ing BCI control, there is specific, targeted control to a specific area of the brain (here, the left motor cortex) in addition to other ongoing control and sensory processes. We wished for our selection of the input matrix B to reflect this richness and also allow for computationally tractable calculations of optimal control, which is diffi- cult for sparse control sets. Therefore, we constructed the input matrix B so as to allow input that was domi- nated by the BCI control site, while maintaining minor contributions from other areas. More specifically, rather than being characterized by binary state values, chan- nels other than the one located over left motor cortex were given a value of approximately 5 × 10−5 at their corresponding diagonal entry in B. See Supplement for the full derivation from [62]. It is important to note that in general the tools from linear controllability theory are not applicable to the functional networks commonly derived from neuroimag- ing data for two reasons. The first reason is that the model which the tools are built upon stipulates a time- dependent propagation of activity along edges; such a propagation is physically true for structural connections derived from white matter, but is not generally true for other types of connections used in network models, such as morphometric similarity or most common functional connectivity measures. While we do not expect that these simple models will fully capture neural dynamics, it is important to explore how much variance these models do explain, even if we expect that amount to be small. The second reason is that the model assumes that in- teractions between nodes "a" and "c" are not due to A central hypothesis in this work is that certain reg- ularized subgraphs are better suited to drive the brain to patterns of activity that are beneficial for BCI con- trol than others. To test this hypothesis, we create tar- get states that reflect these beneficial patterns, based on previous literature. Target states for motor im- agery and attention are obtained for each band indi- vidually from references [20, 23, 24], and can be briefly described as follows: α contralateral motor suppression for motor imagery and parietal suppression for atten- tion, β contralateral motor suppression and ipsilateral motor activation for motor imagery and vertex suppres- sion for attention, and γ contralateral motor activation for motor imagery and motor cortex suppression with frontal and occipital activation for attention (Fig. S9). Channels were divided into lobes using standard mon- tages provided by Brainstorm [61] software (neuroim- age.usc.edu/brainstorm/Tutorials/MontageEditor). The target state of channels in brain regions where we did not have specific hypotheses for their activity were set to zero; the target state of channels with activation were set to 1 and that of channels with deactivation were set to -1. Initial states were set to 0 for all channels. We then calculate the optimal energy (using the optimal con- trol equation described above) required to reach each of these target states to test the hypothesis that subgraphs that support learning will have lower energy requirements than those that do not. Statistical Analyses Much of our analyses involve testing differences in distributions across subjects for different subgraphs the lm.beta package calculated coefficients or sessions, both for phase-randomized and empirical data. We also compare these distributions to subject learning rate defined as the slope of performance over time. For the results displayed in Fig. 2 here in the main manuscript, we used a repeated measures ANOVA to test for the presence of a main effect across conditions given that the distributions of performances were normal (see Fig. S10). In Fig. 3 here in the main manuscript, we sought to predict learning rate with ranked performance loading. After plotting quantile-quantile plots (see Fig. S11-S13) for the learning rate, and each of the performance loadings, it became clear that the lowest loadings were not normally distributed. Therefore, we used a linear model combined with non-parametric test- ing utilizing 5000 permutations (lmPerm package in R https://cran.r-project.org/web/packages/lmPerm). were us- Standardized ing in R (https://cran.r- project.org/web/packages/lm.beta/lm.beta.pdf). We use a Bonferroni correction to control false positive errors due to multiple comparisons across all 6 pre- dictors (α = 0.008). To obtain an estimate of how sensitive our results are to our specific sample, we also plot summary statistics from 500 models obtained from bootstrapping a sample of equal size (N = 60, 3 band and 20 subjects). To examine differences in consistency (Fig. 4 here in the main manuscript), we use a linear model (consistency ∼ band + dataT ype + rank) to test for a main effect of data type (null or empirical), band, and subgraph on consistency (see Fig. S14). We next sought to determine if different subgraphs had consistently different temporal expression for null and empirical data (Fig. 5 here in the main manuscript). We also used a repeated measures ANOVA to test for a main effect of subgraph across bands, and paired t-tests to test for differences amongst individual subgraphs (Fig. S15). Lastly, for the results shown in Fig. 6 here in the main manuscript, we test the relationship between learning rate and optimal control energy differences for several different models. Pearson's correlations were used, given that the data appears normally distributed and has few outliers (see Fig. S16-S19). Data and Code Code for analyses unique to this manuscript are avail- able at github.com/jastiso/netBCI. Code for the NMF al- gorithm and the NMF parameter selection is available at github.com/akhambhati/Echobase/tree/master/Echobase for /Network/Partitioning/Subgraph. at optimal is github.com/jastiso/NetworkControl. neces- sary to reproduce each figure will be made available upon request. Code available analyses control Data 8 RESULTS BCI Learning Performance Broadly, our goal was to examine the properties of dynamic functional connectivity during BCI learning, and to offer a theoretical explanation for why a certain pattern of connectivity would support individual differ- ences in learning performance. We hypothesized that de- composing dynamic functional connectivity into additive N × N subgraphs would reveal unique networks that are well suited to drive the brain to patterns of activity as- sociated with successful BCI control. We use MEG data from 20 healthy adult individuals who learned to control a motor-imagery based BCI over four separate sessions spanning a two week period. Consistent with prior re- ports of this experiment [32], we find a significant im- provement in performance across the four sessions (one- way ANOVA F (3, 57) = 13.8, p = 6.8−7) (Fig. 2). At the conclusion of training, subjects reached a mean per- formance of 68%, which is above chance (approximately 55 - 60%) level for this task [65]. Dynamic patterns of functional connectivity supporting performance To better understand the neural basis of learning per- formance, we detected and studied the accompanying patterns of dynamic functional connectivity. First, we calculated single trial phase-based connectivity in MEG data in three frequency bands: α (7-14 Hz), β (15-25 Hz), and γ (30-45 Hz). We then used non-negative matrix fac- torization (NMF) -- a matrix decomposition method -- to separate the time-varying functional connectivity into a soft partition of additive subgraphs. We found that the selected parameters led to an average of 7.4 subgraphs, with a range of 6 to 9, and that all frequency bands had a decomposition error lower than 0.47 (mean α error = 3.52, mean β error = 0.379, mean γ error = 0.465) (Fig. S2). The error is the Frobenius norm of the squared dif- ference between our observed and estimated connectivity matrices (with dimensions 5152 × 384) and takes values between 0 and 1. For each band, the error value is low, giving us confidence that we have fairly accurately recon- structed relevant neural dynamics. To determine whether any properties of the identified subgraphs were trivially due to preprocessing choices, NMF parameters, or time- invariant autocorrelation in neural activity, we repeated the full decomposition process after permuting the phases of all time series uniformly at random. We found that the statistics of subgraph number and decomposition er- ror were similar for the uniformly phase randomized data, indicating that any differences in subgraph and tempo- ral expression between null and empirical data is not due to the NMF algorithm's inability to find a good decom- position, but rather due to the structure of the chosen decomposition (Fig. S2). We quantified the similarity between each subgraph's temporal expression and the time course of performance, and we refer to this quantity as the subgraph's per- formance loading (Fig. 1). We hypothesized that the ranked performance loading would predict task learn- ing, as operationalized by the slope of performance over time. It is important to note the distinction between per- formance and learning: performance is defined as task accuracy and therefore varies over time, while learning is defined as the linear rate of change in that perfor- mance over the course of the experiment (384 trials over 4 days). We tested whether learning was correlated with the performance loading of subgraphs. Because the min- imum number of subgraphs in a given subject was 6, we decided to investigate the top four highest perfor- mance loading subgraphs, and the smallest and second smallest nonzero loading subgraphs. We found a general trend that the performance loading from high loading subgraphs was negatively associated with learning rate, and the performance loading from low loading subgraphs was positively associated with learning rate (Fig. 3AB). We assessed the statistical significance of these trends and found that only the third highest loading subgraph displayed a performance loading that was significantly correlated with learning rate after Bonferroni correction for multiple comparisons (linear model with permuta- tion tests slope ∼ loading3 + band : p = 0.005). Per- formance loading from uniformly phase randomized sur- rogate data for this subgraph did not predict learning rate (p = 0.292). The direction of the observed effect in the empirical data is notable; subjects with lower load- ing onto high loading subgraphs learned the task better, suggesting that learning is facilitated by a dynamic in- terplay between several subnetworks. It is also notable that the highest loading subgraphs are not the strongest predictors of learning, indicating that the subgraphs that most closely track performance are not the same as the subgraphs that track changes in performance. Spatial properties of dynamic patterns of functional connectivity Next we sought to better understand why the third highest loading subgraph most robustly predicted learn- ing. We hypothesized that because of this subgraph's predictive power across subjects, it might recruit consis- tent brain regions and reflect the involvement of specific cognitive systems across subjects. To evaluate this hy- pothesis, we began by investigating the shared spatial properties of this subgraph in comparison to the others. To identify shared spatial features we grouped subgraphs together by their ranked performance loading, and then quantified how consistent edges were across participants [60] (see Methods). We found that the average consis- tency varied by frequency band, and differed between the empirical and surrogate data, but not across ranked sub- graphs (linear model consistency ∼ band + rank + data : 9 FIG. 3. Performance loading predicts learning. (A) Here we show the p-values for empirical (green) and uniformly phase randomized (grey) data for linear models predicting the slope of performance with ranked performance loading from each frequency band. The black line corresponds to p = 0.05, while the red dashed line corresponds to the Bonferroni corrected α = 0.008. Error bars show the standard error and median of p values from 500 models with bootstrapped samples. (B) The standardized regression coefficients for the same models. Error bars show the standard error and mean of coefficients from 500 models with bootstrapped samples. Fband(2, 17) = 90.36, pband = 9.00×10−10, Fdata(1, 17) = 41.8, pdata = 5.78 × 10−6). The α band had the most consistent edges, followed by the γ band, and then the β band (tαβ = −12.68, pαβ = 4.3 × 10−10, tαγ = −10.41, pαγ = 1.2 × 10−8). In the uniformly phase randomized surrogate data, we observed less consistent subgraphs than those observed in the empirical data (t = −6.47, p = 5.78 × 10−6). These observations support the con- clusion that across the population, despite heterogeneous performance, similar regions interact to support perfor- mance and learning to varying degrees. Anatomically, subgraphs were dominated by connec- tivity in the frontal lobe sensors, with subtle differences in the pattern of connections from the frontal lobe sen- sors to sensors located in other areas of the brain (Fig. 4). To determine which functional edges were most consis- tent in each subgraph and frequency band, we calculated the average consistency over each lobe and motor cortex in both hemispheres (for the same analysis in surrogate data, see Fig. S6). In the α band, the most consis- tent edges on average were located in the left frontal lobe in the highest performance loading subgraph, in the left occipital lobe in the second highest performance load- ing subgraph, between right frontal and right motor in -2.0-1.0-0.50.0Highest2nd H 3rd 4thLowest2nd Llog10(p−value)EmpiricalNull00.2Standardized Coefficientp = 0.05ABHighest2nd H 3rd 4thLowest2nd L-0.2 p = 0.006-1.5-0.4 the third highest performance loading subgraph, and be- tween left frontal lobe and right parietal lobe in the low- est performance loading subgraph. In the β band, the most consistent edges were located between right and left frontal lobe for the highest and second highest per- formance loading subgraph, between left frontal lobe and right motor for the third highest performance loading subgraph, and between left and right frontal lobe for the lowest performance loading subgraph. In the γ band, the most consistent edges were located in the left frontal and right frontal lobes for the highest performance load- ing subgraph, in the left frontal lobe and right motor for the second highest performance loading subgraph, and in left frontal and right frontal lobe for the third highest and lowest performance loading subgraphs. We also note that the most consistent individual edges for each subgraph are still only present in 10-12 individuals, indicating a high amount of individual variability. Collectively, these observations suggest widespread individual variability in the spatial composition of ranked subgraphs, with the most consistent connectivity being located in the frontal lobe during BCI learning. 10 later than all others (paired t-test N = 20, t2H = 10.9, p2H = 1.39 × 10−9; t3H = 7.56, p3H = 3.57 × 10−7; tlow = 8.07, plow = 1.49−7). In the γ band, the highest performance loading subgraph peaked significantly later than the second highest, and lowest loading subgraphs (paired t-test N = 20, t2H = 4.50, p2H = 2.46 × 10−4; tlow = 8.06, plow = 1.49 × 10−7). (Fig. 5). Finally, we asked whether the time of the peak in the third highest performance loading subgraph predicted learn- ing. We did not find a relationship between peak time and learning in any frequency band (Pearson's correla- tion: α : r = 0.005, p = 0.98, β : r = 0.047, p = 0.84, γ : r = −0.21, p = 0.037). To summarize these findings, we note that across participants and especially in the β band, subgraphs that support performance are highly expressed late in learning, when performance tends to be highest. However, subgraphs that support learning do not have consistent peaks across subjects, and each individual's peak does not predict their learning rate, in- dicating that some other feature of these subgraphs must explain their role in learning. Temporal properties of dynamic patterns of connectivity supporting BCI learning via network functional connectivity control theory Explaining dynamic patterns of functional Importantly, subgraphs can be characterized not only by their spatial properties, but also by their temporal expression. We therefore next examined the temporal properties of each subgraph to better understand why the third highest performance loading subgraph most ro- bustly predicted learning. As a summary marker of tem- poral expression, we calculated the total energy of the time series operationalized as the sum of squared values, as well as the time of the peak value of the time series. Across frequency bands, we found no significant depen- dence between energy and subgraph ranking. We did find a significant effect of rank for the peak time of temporal expression obtained from the empirical data (repeated measures ANOVA peak ∼ rank + band : Frank(3, 215) = 6.67, prank = 2.53 × 10−4 but not from the uniformly phase randomized surrogate data (Frank(3, 215) = 1.28, p = 0.282). Overall, peak times are widely distributed across individuals. However we find that across bands, the highest performance loading subgraph has a later peak, which is intuitive since performance is generally increasing over time and these subgraphs most strongly track performance. We then performed post-hoc paired t-tests corrected for multiple comparisons (Bonferroni correction α = 0.006) between the highest performance loading subgraph and all other ranked subgraphs in each band. In the α band, the highest performance loading subgraph only peaked significantly later than the lowest (paired t-test N = 20, tlow = 8.06, plow = 1.49 × 10−7) after Bonfer- roni correction (α = 0.006). In the β band, the high- est performance loading subgraph peaked significantly Lastly we asked how the third highest loading sub- graph could facilitate successful BCI performance, as shown in Fig. 3. Here, we considered an edge -- extracted under penalties of spatial and temporal sparsity -- as a po- tential path for a brain region to affect a change in the ac- tivity of another brain region [66 -- 68]. Assuming the true connectivity structure is sparse, the regularization ap- plied in the NMF algorithm can remove large statistical relationships between regions that are not directly con- nected, but might receive common input from a third re- gion [69, 70] (see Methods for addition discussion, and see Fig. S1A-B for the effect of regularization on the preva- lence of triangles). We hypothesized that the pattern of edges in this subgraph would facilitate brain states, or patterns of activity, that were predictive of BCI literacy. Specifically, we expected that when the brain mirrored the connectivity of the third subgraph, the brain could more easily reach states of sustained motor imagery or sustained attention than when the brain mirrored the connectivity of the lowest performance loading subgraph. We also hypothesized that the magnitude of this differ- ence would be predictive of each subject's learning rate. To test these hypotheses, we used mathematical mod- els from network control theory to quantitatively esti- mate the ease with which the brain can reach a desired pattern of activity given a pattern of connectivity (see Methods and Fig. S1C-D for analyses demonstrating the efficacy of the regularized subgraphs in linearly predict- ing changes in activity). Specifically we calculated the optimal control energy required to reach a target state (either sustained motor imagery or sustained attention) 11 FIG. 4. Spatial distribution of subgraph edges that are consistent across participants. Consistent edges for each frequency band and for each ranked subgraph. Left images show individual edges plotted on a topographical map of the brain. Right images show the mean edge weight over sensors for a given region. We studied 10 regions, including the frontal lobe, temporal lobe, parietal lobe, occipital lobe, and motor cortex in both hemispheres. The weight of the edge corresponds to the number of individual participants for whom the edge was among the 25% strongest for that subgraph. from an initial state when input is applied primarily to the left motor cortex, which was the site of BCI control (Fig. 6A-B). We tested whether the third highest performance load- ing subgraph supported the transition to states of sus- tained motor imagery or sustained attention with smaller energy requirements than other subgraphs that did not support learning in the same way. We chose the lowest performance loading subgraph for comparison because it was the only subgraph with a large positive standard- ized regression coefficient for predicting learning, which contrasts sharply with the large negative coefficient for the third subgraph. For both states (motor imagery and attention), we found no population level differences in energy requirements by the two subgraphs (paired t-test tα = −0.005, pα = 0.565, N = 20, motor imagery: tβ = 1.38, pβ = 0.184, tγ = −1.00, pγ = 0.329. atten- tion: tα = −1.35, pα = 0.193, tβ = −0.344, pβ = 0.735, tγ = −0.937, pγ = 0.360). We next tested whether the magnitude of the difference in energy required by the two subgraphs to reach a given state tracked with learn- ing rate. In the β band, we observed a significant cor- relation between the magnitude of the energy difference to reach attentional states and learning rate over sub- jects (Pearson's correlation coefficient r = 0.560, p = 0.0103, Bonferroni corrected for multiple comparisons 12 (Pearson's correlation r = 0.40, p = 0.077), the second highest with the third highest performance loading sub- graph (Pearson's correlation r = 0.266, p = 0.257), or the second highest with the lowest performance loading subgraph (Pearson's correlation r = −0.072, p = 0.764). This pattern of null results underscores the specificity of our finding. Reliability and specificity of inferences from network control theory Collectively, our findings are consistent with the hy- pothesis that during BCI learning, one subnetwork of neural activity arises, separates from other ongoing pro- cesses, and facilitates sustained attention. An alternative hypothesis is that our results are due to trivial factors re- lated to the magnitude of the attentional state, or could have just as easily been found if we had placed input to a randomly chosen region of the brain, rather than to the left motor cortex which was the actual site of the BCI control. To determine whether these less interest- ing factors could explain our results, we performed the same network control calculation but with a spatially non-overlapping target state, and then -- in a separate simulation -- with a mirrored input region (right motor cortex rather than left motor cortex). We performed the spatial shifting by ordering the nodes anatomically (to preserve spatial contiguity), and then circular shifting the attention target state by random number between 1 and N−1. For 500 circularly shifted states, only 3 (0.6%) had a correlation value equal to or stronger than the one observed (Fig. S7). Furthermore, we found no significant relationship between learning rate and the difference in energy required by the two subgraphs to reach the true attention state when input was applied to the right motor cortex instead of the left motor cortex (Pearson's corre- lation t = 0.711, p = 0.313). Together, these two findings suggest that the relationship identified is specific to BCI control. Finally, we assessed the robustness of our results to choices in modeling parameters. First we performed the computational modeling with two different sets of control parameter values (see Supplement). In both cases, the significant relationship remained between learning rate and the difference in energy required by the two sub- graphs to reach the attentional state (set one Pearson's correlation coefficient r = 0.476, p = 0.0338; set two Pearson's correlation coefficient r = 0.514, p = 0.0204). Second, since our target states were defined from prior literature, there was some flexibility in stipulating fea- tures of those states. To ensure that our results were not unduly influenced by these choices, we tested whether ideologically similar states would provide similar results. Namely, we assessed (i) the impact of varying the mag- nitude of (de)activation by changing (-)1 to (-)2, (ii) the impact of the neutral state by changing 0 to 1, and (iii) the impact of negative states by changing -1, 0 and 1 to FIG. 5. Temporal expression of ranked subgraphs. The peak temporal expression for every subject (black data point), for each frequency band (indicated by color) and for each subgraph (ordered vertically). Violin plots show the density distribution of all subjects' peaks. The median is marked with a solid line through the violin plot. across frequency bands; Fig. 6). Notably, the relation- ship remained significant when controlling for subgraph density (linear model slope ∼ energy dif f erence + density dif f erence: tenergy = 2.68, penergy = 0.0158, tdensity = −0.266, pdensity = 0.794). When using sub- graphs derived from the uniformly phase randomized sur- rogate data, the relationship was not observed (Pearson's correlation r = −0.0568, p = 0.819). We next asked which subgraph contributed most to this effect. We found no significant relationship between learning rate and the energy required to reach the attentional state by the third highest performance loading subgraph (Pear- son's correlation r = −0.389, p = 0.702) or by the low- est performance loading subgraph (Pearson's correlation r = 0.227, p = 0.335). This finding suggests that learn- ing rate depends on the relative differences between sub- graphs, rather than the energy conserving architecture of one alone. As a final test of specificity, we assessed whether this difference was selective to the third highest and lowest performance loading subgraph. We found no significant relationship when testing the difference of the highest with the third highest performance loading sub- graph (Pearson's correlation r = −0.554, p = 0.586), the highest with the lowest performance loading subgraph 0.250.50.751Lowest2HHighestNormalized Peak03HNormalized TimeBand:AlphaBetaGammaRank(last trial)((cid:31)rst trial)0 13 FIG. 6. Separation of the ability to modulate attention predicts learning. Different patterns of connections will facilitate transitions to different patterns of brain activity. We hypothesize that the ease with which connections in certain regularized subgraphs facilitate transitions to patterns of activity that support either motor imagery (A) or attention (B) will predict learning rate. We use network control theory to test this hypothesis. We model how much energy (u(t)) is required to navigate through state space from some initial pattern of activity x(0) to a final pattern of activity x(T ). Some networks (e.g., the brown network in panel A) will require very little energy (schematized here with a smaller, solid colored arrow) to reach patterns that support motor imagery, while other networks (e.g., the pink network in panel B ) will have small energy requirement to reach patterns of activity that support attention. (C) The relationship between learning rate and the difference in energy required to reach the attention state when the underlying network takes the form of the lowest versus third highest performance loading subgraphs for empirical data (green) and uniformly phase randomized surrogate data (grey). (D) The relationship between the learning rate and the energy required to reach the attention state when the underlying network takes the form of the lowest performance loading subgraph, or when the underlying network takes the form of the third highest performance loading subgraph. 1, 2, and 3. We found a consistent relationship between learning rate and the difference in energy required by the two subgraphs to reach the attentional state when we changed the magnitude of activation/deactivation (Pear- son's correlation coefficient r = 0.560, p = 0.0103), as well as when we changed the neutral state (Pearson's correlation coefficient r = 0.520, p = 0.0188). However, we found no significant relationship when removing neg- ative states (Pearson's correlation coefficient r = 0.350, p = 0.130), indicating that this result is dependent on our choice to operationalize deactivation as a negative state value. After performing these robustness checks, we con- clude that a selective separation of the third highest and lowest performance loading subgraphs impacts their abil- ity to drive the brain to patterns of sustained attention in the β band in the context of BCI control. This result is robust to most of our parameter choices, is selective for biologically observed states, and is not observed in surrogate data. DISCUSSION In this work, we use a minimally constrained decom- position of dynamic functional connectivity during BCI learning to investigate which groups of phase locked brain regions (subgraphs) support BCI control. The per- formance loading onto these subgraphs favors the the- ory that dynamic involvement of several subgraphs dur- ing learning supports successful control, rather than ex- tremely strong expression of a single subgraph. Addi- tionally, we find a unique role for the third highest load- ing subgraph in predicting learning at the population level. This result shows that learning is not simply pre- dicted by the subset of edges that has the most simi- lar temporal expression, but rather that a subnetwork with a middling range of similarity has the strongest re- lationship with performance improvement. While the spatiotemporal distribution of this subgraph was vari- able across individuals, we did observe some consistencies at the group level. Spatially, the third highest loading subgraph showed strong edges between left frontal and right motor cortices for low frequencies, and left frontal and left motor cortices for the γ band. Lower frequen- cies showed stronger connectivity to the ipsilateral (to imagined movement) motor cortex, suggesting a possible role in suppression for selective control. This subgraph also showed the highest expression earlier than the other ranked subgraphs we investigated, perhaps linking it to the transition from volitional to automatic control. We next wished to posit a theory of how these sub- graphs fit with previously identified neural processes im- portant for learning, despite their heterogeneity across subjects. After quantifying the extent to which NMF reg- ularization removed potentially redundant relationships between regions (Fig. S1A-B), we suggested that the regularized pattern of statistical relationships identified in this subgraph could comprise an avenue through which brain activity could be modulated via cognitive control or external input. We then hypothesized that these net- works would be better suited to modulate activity in either regions implicated in attention or in motor im- agery than other subgraphs, and further that individuals whose networks better modulated activity in these re- gions would display greater task learning [18]. We chose to operationalize the "ease of modulation" with a met- ric from network control theory called optimal control energy. Optimal control energy quantifies the minimum input needed to drive the brain from an initial pattern of activity to a final pattern of activity, while also assuring that the pattern of activity stays close to the target state at every point in time. This last constraint assures that we do not pass through biologically unfeasible patterns of activity to reach our desired pattern. The notion of optimal control energy that we use here assumes a par- ticular linear model of how neural dynamics change given potential avenues of communication between regions. Im- portantly, in the supplement (Fig. S1C-D) we show that our subgraphs predict empirical brain state changes ac- cording to this model, and that the contribution of each subgraph to empirical changes in brain state is related to its temporal expression. Using this model, we did not find any population differences in optimal control energy when the simulation was enacted on the third highest performance loading subgraph compared to the lowest performance loading subgraph. However, we did find that the magnitude of this difference predicted learning in individual subjects. This result was specific to the β band and to brain regions implicated in attention. Crit- ically, the relation to learning could not be explained by the energy of either subgraph alone, was not present in surrogate data derived from a uniformly phase randomize null model, and was robust to parameter choices. Over- all, the observations support our hypothesis that in the β band the subgraphs we identified that support learn- ing are well suited to modulate activity in brain regions associated with attention. 14 A delicate balance of interactions is required for BCI learning Our initial analysis explored the relationship between performance loading and learning. It is important to note the behavioral difference between performance and learning: we use the term performance to refer to task accuracy over time, whereas we use the term learning to refer to how well a subject is able to increase that accu- racy. With that distinction in mind, we aimed to better understand how subgraphs that vary similarly to perfor- mance (those with high performance loading) relate to learning. We found that the subgraph with the third highest performance loading best predicted learning and that a narrow distribution of performance loading across all subgraphs was associated with better learning. To- gether, these two observations are in line with previous research in motor and spatial learning, which shows that some brain structures display differential activity during learning that is independent of performance [71, 72]. Our work adds to this literature by demonstrating that in ad- dition to targeted differences in individual brain regions or networks, a minimally constrained decomposition of dynamic functional connectivity across the whole brain reveals that separable processes are most associated with performance and with learning. Additionally, we find that BCI learning is not pre- dicted simply by the processes most strongly associated with performance and learning individually, but by a dis- tributed loading across many different subgraphs. This notion is supported by the modestly predictive role of standard deviation of loading predicting learning, and also by the sign of beta value predictors for ranked sub- graphs. Generally, subgraphs with higher ranked load- ing were negative predictors, while subgraphs with lower ranked loading were positive predictors. A wealth of whole brain connectivity analyses have similarly shown that the interaction between systems is an important component of skill learning specifically, and other do- mains of learning more generally [30, 31, 73, 74]. While we observed marked interactions between many regions, the majority were located in the frontal lobe for all fre- quency bands. Previous work has also demonstrated changes in frontal-motor [75, 76] and fronto-parietal [77] connectivity during motor skill learning. In BCI learn- ing specifically, the strength of white matter connectivity between frontal and occipital regions predicts control of motor imagery based BCIs [78]. Additionally, analyses of this same experiment have shown task related changes in functional connectivity were spatially diffuse, and found in frontal, temporal, and occipital regions in the α band [32], and were strongest in frontal, motor, central, and parietal regions in the β band. Our results add to these findings by demonstrating that the most consistent re- gions that covary in their functional connectivity are in- teractions between the frontal lobe and other regions. Our work shows that broad motifs like the dynamic inte- gration of multiple systems (including cognitive systems involving the frontal lobe) found in other types of learn- ing are also important for BCI learning. Additionally, we add to previous work on BCI learning specifically by quantifying the structure of covarying subgraphs of con- nectivity. BCI learning is heterogenous across individuals We find population level consistencies in spatial and temporal properties of ranked subgraphs despite hav- ing no constraint to assure consistency across individ- uals. However, we also note that there is a high degree of variability in both of these measures. The variabil- ity is mirrored in the subjects' performance, with final performances varying from 38.1 % to 89.3 %. Our ob- servations are in line with previous literature demon- strating variability in subjects' performance and learn- ing for psychological, cognitive, and neurological predic- tors [18, 79]. Such pervasive and marked individual dif- ferences presents a challenge for the use of BCIs clini- cally [80]. To address this challenge, researchers have explored ways to optimize BCI features and algorithms for neurofeedback itself [81, 82] and to identify selection criteria for BCI based therapies [57, 79]. The results of our study support the idea that different individuals will have slightly different neural correlates of both perfor- mance and learning based on a variety of features such as demographics [83], spatial manipulation skills [84], re- lationship with the technology [85], and attention span [23, 86]. Our findings also highlight the importance of studying models fit to each individual when searching for selection criteria for BCI therapies. Here, despite tem- poral and edge level heterogeneity, our minimally con- strained, individual specific method of brain connectiv- ity decomposition revealed a robust predictor of learning with a theoretical role that aligns well with previous liter- ature. Further development and expansion of this model to incorporate resting state neuroimaging data and other physiological predictors could be a promising direction for selection of candidates for BCI therapies before train- ing. Role of beta oscillations in BCI learning Prominent theories describing the neural processes that give rise to cognition and shape our behavior often involve integration of complex multimodal information using a combination of top-down predictions (built from prior experience) and bottomup, sensory-driven represen- tations of the dynamic world around us [87 -- 89]. These generalized frameworks, in turn, require the precise co- ordination of ensemble neural activity both within and between brain regions. Several theoretical approaches have examined how these two scales of functional activ- ity may harmonize to produce the desired behavior [90], and empirical research has shown that there is consistent cross-talk between these scales [91]. Within human neu- 15 roimaging work, synchronous oscillations have been crit- ical to understanding this complex coordination, where cortico-cortical propagation delays and membrane poten- tials give rise to observed oscillatory activity in the brain [92, 93]. Here, we study the time varying connectivity within α, β, and γ bands. Much like how specialized functions arise from different brain regions, different nar- rowband oscillations have been implicated in diverse but specialized processes, where some generalizable theories suggest a role for α in disengagement of task irrelevant areas or a lack of sensory processing [94], β in sustain- ing the current cognitive state [95] and γ in task active local cortical computation [96]. Specifically in the con- text of motor imagery based BCIs, α and β bands have prominent signatures in motor imagery [97]. Our results show that only the β bands functional connectivity is well suited to modulate patterns of activity that support sus- tained attention (not motor imagery), which is a critical process for BCI control. While our results are in line with generalized theories on the role of oscillations in cogni- tion, the specificity of the β band in our results extends classic studies that discuss the role of this oscillation in attention [98] and in maintaining the current cognitive state [95]. Our results suggest that this maintenance, a consistent control (or attention to) internally generated activity, may play a crucial role in longterm BCI use. Methodological Considerations NMF Non-negative matrix factorization is a machine learning technique for separating, in our case, a multi- modal configuration matrix into a soft-partition of sub- graphs with time-varying expression. This process has several advantages, such as being able to link behav- ioral and neural data, and creating a quantification of mesoscale structure where brain regions can participate in multiple functional groups. Nevertheless, the method also faces several limitations that are common to other large-scale machine learning techniques. NMF yields a low rank approximation of a large configuration matrix, and can sometimes be rank deficient for large number of subgraphs, for very large datasets, or for datasets with high covariance. Because of this sensitivity, we were not able to test our data against independently phase ran- domized null models. Spatial Resolution We chose to complete our analyses in sensor, rather than source space. This choice limits the anatomical resolution of our data, and therefore the specificity of the claims that we can make about the spa- tial distribution of the regions involved. However, source reconstruction requires many parameter choices and has potentially confounding effects on estimates of functional connectivity [99 -- 101]. Additionally, we were not inter- ested in finer resolution distribution of the identified sub- graphs, but more in the process of identifying them, in validating the hypothesis that features of these subgraphs predicted learning, and in their theoretical functions. Optimal Control We chose to use tools from network control theory to quantify the ease with which each net- work can modulate brain activity. Network control the- ory relies on several assumptions that should be consid- ered when interpreting these results. First, the model of dynamics that we employ is linear and noise free, unlike the brain [102], but has proven useful in gaining intuitions about the behavior of nonlinear systems [103, 104]. How- ever, we still sought to quantify the ability of this linear model to explain empirical changes in brain state. Specif- ically, we asked two questions: (1) do the regularized subgraphs used in our analyses have the ability to predict state transitions, and do they do so better than randomly rewired networks, and (2) is the contribution of each sub- graph to explaining a given state transition proportional to its temporal expression, and is it more proportional than a different subgraph's temporal expression? To eval- uate these questions, we generated brain states for every trial (band specific power at each channel) and simulated Eq. 5 (see Supplement). Regarding the similarity of pre- dicted and empirical state transitions, we find modest correlation values (mean Pearson's r = 0.25) that are significantly greater than the correlations observed from randomized networks. Similarly for our second question, we found small but positive correlations between the con- tribution of each subgraph to a given transition and its temporal expression (mean Pearson's r = 0.03), which was also significantly greater than correlations to tempo- ral expression from mismatched subgraphs. While it is unsurprising that our linear model did not fully capture neural dynamics across a three second trial, it is worth considering extensions that can maximize this similar- ity for future analyses investigating how connections be- tween regions facilitate changes to activity. One option is to use effective connectivity -- such as autoregressive models [69, 105] -- that solve for a network of connec- tions that best predicts the evolution of brain states in time. However, effective connectivity matrices are of- ten sparse, and therefore not well suited to NMF matrix decomposition used in the present work. Alternatively, one could use non-linear models of dynamics [106] and non-linear control theory [107] to capture a wider range of dynamic behaviors, although non-linear control does not currently support the same scope of tools available for linear control theory. Lastly, future work could use functional approximation[108] in order to identify a set of simple basis functions that well approximate the data. If a sparse approximation can be found, it supports the idea that the underlying non-linear dynamics can be cap- tured with linear combinations of these basis functions, and therefore are suitable to be modeled with simplified linear models. Additionally, network control is typically applied to time invariant, structural connections that have a clear role as an avenue along which brain activity can prop- agate. Here we used functional connectivity (weighted phase locking) which is a statistical relationship that (1) does not imply the presence of a physical connection and 16 (2) is not time invariant. Due to (1), our original func- tional connectivity matrix can have large values between two regions that are not directly connected, but might both connect to the same region. This situation would lead to a triangle composed of three connections in a functional connectivity matrix where in reality there are only two connections. However, the regularization ap- plied by the NMF algorithm mitigates this concern in a manner that is similar to the regularization applied in effective connectivity metrics [69, 70]. We also explicitly quantify the effect of regularization on triangles in our subgraphs and find a dramatic reduction from the origi- nal functional connectivity (Fig. S1A-B). This quantifi- cation, along with the two validations discussed above, show that our model is a suitable way to evaluate the role of regularized subgraphs in modulating different patterns of activity. In relation to (2), we note that functional connectivity in not time-invariant, unlike the state ma- trix more commonly employed in linear control models. However, it is important to note that NMF identifies sub- graphs that are separable from their temporal expression, and that we expect that the hypothesized role in control would only be prominent when the subgraph was highly expressed. Conclusion In conclusion, we use a minimally constrained method of matrix decomposition that is specific to each human participant to investigate the dynamic neural networks that support BCI learning. We find that the subgraphs that most tightly mirror performance are not the same subgraphs that most strongly support learning. Addi- tionally, we find that the interaction between many differ- ent neural processes is important for BCI learning. While the subgraphs identified are heterogeneous (as is subject performance), we find consistent involvement of frontal and motor cortices in subgraphs that support learning. We also observe differential temporal expression amongst subgraphs, and perhaps most notably that the subgraphs that vary more similarly with performance reach their highest expression later in learning. Lastly, we test the hypothesis that subgraphs that support learning are bet- ter suited to modulate activity in brain regions important for attention than other subgraphs. We find evidence to support this hypothesis in the β band specifically, ulti- mately suggesting that the separation of processes for maintaining attention is important for successful BCI learning. Our results align with prior work from dynamic functional connectivity in other types of skill learning, and also highlight a method for identifying individual predictors of successful BCI control with theoretical sup- port. 17 ACKNOWLEDGEMENTS We would like to thank Ankit N. Khambhati for help- ful discussions regarding the application of NMF to BCI learning, and Marcelo G. Mattar for helpful dis- cussions regarding joint decomposition of brain and be- havior. We would also thank Pragya Srivastava, Ja- son Kim, Lia Papadopoulos, and Eli Cornblath for their helpful comments on the manuscript. D.S.B. and J.S. acknowledge support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foun- dation, the ISI Foundation, the Paul Allen Founda- tion, the Army Research Laboratory (W911NF-10-2- 0022), the Army Research Office (Bassett-W911NF-14- 1-0679, Grafton-W911NF-16-1-0474, DCIST- W911NF- 17-2-0181), the Office of Naval Research, the National Institute of Mental Health (2-R01-DC-009209-11, R01 MH112847, R01-MH107235, R21-M MH-106799), the National Institute of Child Health and Human Devel- opment (1R01HD086888-01), National Institute of Neu- rological Disorders and Stroke (R01 NS099348), the Na- tional Science Foundation (BCS-1441502, BCS-1430087, NSF PHY-1554488 and BCS-1631550), and French pro- gram Investissements davenir ANR-10-IAIHU-06; AN- RNIH CRCNS ANR-15-NEUC-0006-02. The content is solely the responsibility of the authors and does not nec- essarily represent the official views of any of the funding agencies. 4 [1] R. Sitaram, T. Ros, L. Stoeckel, S. Haller, F. Scharnowski, J. Lewis-Peacock, N. Weiskopf, M. L. Blefari, M. Rana, E. Oblak, N. Birbaumer, and J. Sulzer, Nature Reviews Neuroscience 18, 86 (2017). (1969), [2] E. Science 163, E. Fetz, 955 http://science.sciencemag.org/content/163/3870/955.full.pdf. Biological Psychology 89, 80 (2012). [20] A. Bamdadian, C. Guan, K. K. Ang, and J. Xu, Journal of Neuroscience Methods 235, 138 (2014). [21] A. Guillot, C. Collet, V. A. Nguyen, F. Malouin, and J. Doyon, NeuroImage 41, 1471 C. Richards, (2008). [3] M. D. Sacchet, J. Mellinger, R. Sitaram, C. Braun, N. Birbaumer, and E. Fetz, Frontiers in Neuroscience 6, 1 (2012), arXiv:15334406. [4] B. Graimann, B. Allison, and G. Pfurtscheller, , 1 (2010), arXiv:arXiv:1011.1669v3. [5] K. A. Moxon and G. Foffani, Neuron 86, 55 (2015). [6] D. S. Bassett and O. Sporns, Nature Neuroscience 20, 353 (2017), arXiv:0106096v1 [arXiv:cond-mat]. [7] M. Reiner, R. Rozengurt, and A. Barnea, Biological Psychology 95, 45 (2014). [22] S. Halder, D. Agorastos, R. Veit, E. M. Hammer, S. Lee, B. Varkuti, M. Bogdan, W. Rosenstiel, N. Bir- baumer, and A. Kubler, NeuroImage 55, 1779 (2011), arXiv:j.neuroimage.2011.01.021. [10.1016]. [23] M. Grosse-Wentrup, B. Scholkopf, and J. Hill, (2011), 10.1016/j.neuroimage.2010.04.265. [24] J. Frey, C. Muhl, F. Lotte, and M. Hachet, Proceedings of the International Conference on Physiological Com- puting Systems , 2019 (2013), arXiv:1311.2222. [25] F. De Vico Fallani and D. S. Bassett, Physics of Life [8] T. Ros, M. A. M. Munneke, L. A. Parkinson, and J. H. Reviews 1, 6 (2018), arXiv:1807.05616. Gruzelier, Biological Psychology 95, 54 (2014). [9] V. P. Buch, A. G. Richardson, C. Brandon, J. Stiso, M. N. Khattak, D. S. Bassett, and T. H. Lucas, Fron- tiers in Neuroscience 12, 790 (2018). [10] S. Shahid, R. Kumar Sinha, and G. Prasad, (2010), 10.1186/1471-2202-11-S1-P127. [11] T. Ros, B. J. Baars, R. A. Lanius, and P. Vuilleumier, Frontiers in Human Neuroscience 8 (2014), 10.3389/fn- hum.2014.01008. [12] N. Lofthouse, L. E. Arnold, S. Hersch, E. Hurt, and (2012), R. DeBeus, Journal of Attention Disorders 10.1177/1087054711427530. [13] H. Gevensleben, B. Holl, B. Albrecht, D. Schlamp, O. Kratz, P. Studer, A. Rothenberger, G. H. Moll, and H. Heinrich, 10.1007/s00787-010-0109-5. [14] R. T. Thibault, M. Lifshitz, and A. Raz, Cortex 74, 247 (2016). [15] M. Hamedi, S. H. Salleh, and A. M. Noor, Neural Com- putation 28, 999 (2016), arXiv:1309.2848v1. [16] M. Ahn and S. C. Jun, Journal of Neuroscience Methods 243, 103 (2015). [17] E. A. Curran and M. J. Stokes, Brain and Cognition 51, 326 (2003). [18] C. Jeunet, B. N'kaoua, and F. Lotte, Progress in Brain Research 228 (2016), 10.1016/bs.pbr.2016.04.002. [19] E. M. Hammer, S. Halder, B. Blankertz, C. Sannelli, T. Dickhaus, S. Kleih, K. R. Muller, and A. Kubler, [26] M. E. J. Newman, Networks: An Introduction (2010). [27] E. Bullmore and O. Sporns, Nature Neuroscience (2009), 10.1038/nrn2575. [28] D. S. Bassett, P. Zurn, and J. I. Gold, Nature Reviews Neuroscience , 1 (2018). [29] F. De Vico Fallani, J. Richiardi, M. Chavez, and S. Achard, Philosophical Transactions of the Royal Society B: Biological Sciences 369 (2014), 10.1098/rstb.2013.0521. [30] D. S. Bassett, M. Yang, N. F. Wymbs, and S. T. Grafton, Nature Neuroscience 18, 744 (2015), arXiv:1403.6034. [31] D. S. Bassett and M. G. Mattar, Trends in Cognitive Sciences 21, 250 (2017). [32] M.-C. Corsi, M. Chavez, D. Schwartz, N. George, L. Hugueville, A. E. Khan, S. Dupont, D. S. Bassett, and F. De Vico Fallani, (2018), 10.1101/487074. [33] A. N. Khambhati, A. E. Sizemore, R. F. Betzel, and D. S. Bassett, "Modeling and interpreting mesoscale network dynamics," (2018). [34] M. Fukushima, R. F. Betzel, Y. He, M. A. de Reus, M. P. van den Heuvel, X. N. Zuo, and O. Sporns, "Fluctuations between high- and low-modularity topol- ogy in time-resolved functional connectivity," (2018), arXiv:1511.06427. [35] D. S. Bassett, N. F. Wymbs, M. A. Porter, P. J. Mucha, and S. T. Grafton, Proceedings of J. M. Carlson, the National Academy of Sciences 108, 7641 (2010), arXiv:1010.3775. [36] M. Pedersen, A. Zalesky, A. Omidvarnia, and G. D. [57] C. Jeunet, B. Nkaoua, S. Subramanian, M. Ha- and F. Lotte, PLoS ONE 10, 1 (2015), chet, arXiv:journal.pone.0143962 [10.1371]. Jackson, bioRxiv (2018), 10.1073/pnas.1814785115. [58] J. Kim, Y. He, and H. Park, Journal of Global Opti- 18 [37] J. M. Shine, M. Breakspear, P. T. Bell, K. E. Martens, R. Shine, O. Koyejo, O. Sporns, and R. A. Poldrack, NATuRE NEuROSciENcE -- 22, 10.1038/s41593-018- 0312-0. [38] U. Braun, A. Schafer, D. S. Bassett, F. Rausch, J. I. Schweiger, E. Bilek, S. Erk, N. Romanczuk- Seiferth, O. Grimm, L. S. Geiger, L. Haddad, K. Otto, S. Mohnke, A. Heinz, M. Zink, H. Walter, E. Schwarz, A. Meyer-Lindenberg, and H. Tost, Proceedings of the National Academy of Sciences 113, 12568 (2016). [39] R. T. Gerraty, J. Y. Davidow, K. Foerde, A. Galvan, D. S. Bassett, and D. Shohamy, J Neurosci 2084, 17 (2018). [40] A. N. Khambhati, M. G. Mattar, N. F. Wymbs, S. T. and D. S. Bassett, NeuroImage 166, 385 Grafton, (2018). [41] L. R. Chai, A. N. Khambhati, R. Ciric, T. M. Moore, R. C. Gur, R. E. Gur, T. D. Satterthwaite, and D. S. Bassett, 10.1162/netn a 00001. [42] A. N. Khambhati, D. S. Bassett, B. S. Oommen, S. H. Chen, T. H. Lucas, K. A. Davis, and B. Litt, Eneuro 4, ENEURO.0091 (2017). [43] J.-L. Chen, T. Ros, and J. H. Gruzelier, Human Brain Mapping 34, 852 (2013). [44] R. Monti, R. Lorenz, P. Hellyer, R. Leech, C. Anag- nostopoulos, and G. Montana, Graph embeddings of dynamic functional connectivity reveal discriminative patterns of task engagement in HCP data, Tech. Rep. (2015) arXiv:1506.05219v1. [45] D. D. Lee and H. S. Seung, Nature 401, 788 (1999), arXiv:arXiv:1408.1149. [46] S. Gu, R. F. Betzel, M. G. Mattar, M. Cieslak, P. R. Delio, S. T. Grafton, F. Pasqualetti, and D. S. Bassett, NeuroImage 148, 305 (2017), arXiv:1607.01706. [47] G. Schalk, D. J. McFarland, T. Hinterberger, N. Bir- baumer, and J. R. Wolpaw, IEEE Transactions on Biomedical Engineering 51, 1034 (2004), arXiv:1741- 2560/2/4/008 [10.1088]. [48] C. T. Butts, Science 325, 414 (2009), arXiv:arXiv:1010.0725v1. [49] P. Garc´es, D. L´opez-Sanz, F. Maest´u, and E. Pereda, 10.3390/s17122926. [50] M. Vinck, R. Oostenveld, M. Van Wingerden, F. Battaglia, and C. M. Pennartz, NeuroImage 55, 1548 (2011), arXiv:0006269v1 [hep-ph]. [51] J.-P. Lachaux, A. Lutz, D. Rudrauf, D. Cosmelli, M. Le Van Quyen, J. Martinerie, and F. Varela, Estimating the time-course of coherence between single-trial brain signals: an introduction to wavelet coherence, Tech. Rep. (2002). [52] S. Heitmann and M. Breakspear, Network Neuroscience , 1 (2018), arXiv:arXiv:1404.2263v1. [53] J. Theiler, Testing for nonlinearity in timeseries: the method of surrogate data, Tech. Rep. [54] D. J. Bartholomew, International Encyclopedia of Edu- cation 2, 374 (2010). [55] P. Comon, P. Comon, I. Component, and P. Comon, Signal Processing 1684, 287 (2015). [56] J. Kim and H. Park, SIAM J. Sci. Comput, Tech. Rep. 1 (2011). mization 58, 285 (2014). [59] D. Greene, G. Cagney, N. Krogan, and P. Cunningham, Bioinformatics 24, 1722 (2008). [60] J. A. Roberts, A. Perry, G. Roberts, P. B. Mitchell, and M. Breakspear, NeuroImage 145, 118 (2017). [61] F. Tadel, S. Baillet, J. C. Mosher, D. Pantazis, and R. M. Leahy, Intell. Neuroscience 2011, 8:1 (2011). [62] S. Gu, R. F. Betzel, M. G. Mattar, M. Cieslak, P. R. Delio, S. T. Grafton, F. Pasqualetti, and D. S. Bassett, , 1 (2016), arXiv:1607.01706. [63] R. F. Betzel, S. Gu, J. D. Medaglia, F. Pasqualetti, and D. S. Bassett, Scientific Reports 6, 1 (2016), arXiv:1603.05261. [64] F. Pasqualetti, S. Zampieri, and F. Bullo, IEEE Trans- actions on Control of Network Systems 1, 40 (2014). [65] G. R. Muller-Putz, R. Scherer, C. Brunner, R. Leeb, and G. Pfurtscheller, Ijbem.Org 10, 52 (2008). [66] M. D. Fox, R. L. Buckner, M. P. White, M. D. Gre- icius, and A. Pascual-Leone, Biological Psychiatry 72, 595 (2012). [67] M. D. Fox, H. Liu, and A. Pascual-Leone, NeuroImage 66, 151 (2013). [69] Y. Shen, B. Baingana, [68] F. Ferreri, F. Vecchio, D. Ponzo, P. Pasqualetti, and P. M. Rossini, Human Brain Mapping 35, 1969 (2014). and G. B. Giannakis, IEEE TRANSACTIONS ON MEDICAL IMAGING, Tech. Rep. (2016) arXiv:1610.06551v1. [70] A. Das, A. L. Sampson, C. Lainscsek, L. Muller, W. Lin, J. C. Doyle, S. S. Cash, E. Halgren, and T. J. Sejnowski, Neural Computation, Tech. Rep. 3 (2017) arXiv:1309.2848v1. [71] A. L. Shelton and J. D. E. Gabrieli, American Psycho- logical Association 18, 442 (2004). [72] A. Purushotham, R. D. Seidler, K. U&gbreve;urbil, J. Ashe, S.-G. Kim, and D. Willingham, Science, Tech. Rep. 5575 (2002). [73] N. Altman and M. Krzywinski, Nature Methods 14, 545 (2017), arXiv:NIHMS150003. [74] A. V. Mantzaris, D. S. Bassett, N. F. Wymbs, E. Estrada, M. A. Porter, P. J. Mucha, S. T. Grafton, and D. J. Higham, Journal of Complex Networks 1, 83 (2013). [75] L. Ma, B. Wang, S. Narayana, E. Hazeltine, X. Chen, D. A. Robin, P. T. Fox, and J. Xiong, Brain Research (2009), 10.1016/j.brainres.2009.12.073. [76] D. R. Leff, A. Darzi, K. Shetty, G.-Z. Yang, and J. Andreu-Perez, Brain Connectivity 6, 375 (2016). [77] C. H. Janice Lin, M. C. Chiang, B. J. Knowlton, M. Ia- coboni, P. Udompholkul, and A. D. Wu, Human Brain Mapping 34, 1542 (2013). [78] R. Sitaram, S. Halder, N. Birbaumer, M. Bog- and A. Kubler, (2013), dan, W. Rosenstiel, B. Varkuti, Frontiers arXiv:fnhum.2013.00105 [10.3389]. in Human Neuroscience 7, 1 [79] S. Halder, E. M. Hammer, A. Kubler, B. Blankertz, C. Sannelli, T. Dickhaus, G. Curio, and K.-R. Muller, NeuroImage 51, 1303 (2010). [80] C. Brunner, V. Kaiser, B. Z. Allison, G. Pfurtscheller, G. R. Muller-Putz, and C. Neuper, Journal of Neural Engineering 7, 026007 (2010). 19 [81] C. Vidaurre, C. Sannelli, K. R. Muller, and B. Blankertz, Journal of Neural Engineering 8 (2011), 10.1088/1741-2560/8/2/025009. [82] A. Kubler, M. Tangermann, K.-R. Muller, J. Hohne, F. Cincotti, R. Rupp, J. d. R. Millan, D. Mattia, R. Leeb, and G. Muller-Putz, Proceedings of the IEEE 103, 926 (2015). [83] J. Schumacher, C. Jeunet, and F. Lotte, in 2015 IEEE International Conference on Systems, Man, and Cyber- netics (2015) pp. 3169 -- 3174. [84] A. Vuckovic and B. A. Osuagwu, Clinical Neurophysi- ology 124, 1586 (2013). [85] M. J. Brosnan, Journal of Computer Assisted Learning 14, 223 (1998). [86] M. Grosse-Wentrup and B. Scholkopf, Journal of Neural Engineering 9 (2012), 10.1088/1741-2560/9/4/046001. [87] O. Doehrmann and M. J. Naumer, Brain Research 1242, 136 (2008). [88] P. Kok, J. F. M. Jehee, and F. P. de Lange, Neuron 75, 265 (2012). [89] D. Talsma, Frontiers in Integrative Neuroscience 09, 1 (2015). [90] S. L. Bressler and J. A. Kelso, Trends in cognitive sci- ences 5, 26 (2001). [91] G. Buzs´aki, Nature Neuroscience 7, 446 (2004), arXiv:arXiv:1608.08828v1. [92] A. M. Bastos, W. M. Usrey, R. A. Adams, G. R. Man- gun, P. Fries, and K. J. Friston, Neuron 76, 695 (2012). [93] W. Singer, A. K. Engel, and P. Fries, Nature Reviews. [96] P. Fries, Annual Review of Neuroscience 32, 209 (2009). [97] D. J. McFarland, L. A. Miner, T. M. Vaughan, and J. R. Wolpaw, Brain Topography, Tech. Rep. 3 (2000). [98] G. Pfurtscheller A'b', C. Neuper, D. Flotzinger, and M. Pregenzer, Electroencephalography and clinical Neu- rophysiology, Tech. Rep. (1997). [99] M. J. Brookes, J. R. Hale, J. M. Zumer, C. M. Steven- son, S. T. Francis, G. R. Barnes, J. P. Owen, P. G. Mor- ris, and S. S. Nagarajan, NeuroImage 56, 1082 (2011). [100] A. Hillebrand, G. R. Barnes, J. L. Bosboom, H. W. Berendse, and C. J. Stam, NeuroImage 59, 3909 (2012). [101] G. L. Colclough, M. W. Woolrich, P. K. Tewarie, M. J. Brookes, A. J. Quinn, and S. M. Smith, NeuroImage 138, 284 (2016). [102] S. Gu, F. Pasqualetti, M. Cieslak, Q. K. Telesford, A. B. Yu, A. E. Kahn, J. D. Medaglia, J. M. Vettel, M. B. Miller, S. T. Grafton, and D. S. Bassett, Nature Com- munications 6, 1 (2015), arXiv:1406.5197. [103] S. F. Muldoon, F. Pasqualetti, S. Gu, M. Cieslak, S. T. Grafton, J. M. Vettel, and D. S. Bassett, PLoS Computational Biology 12 (2016), 10.1371/jour- nal.pcbi.1005076, arXiv:1601.00987. [104] C. J. Honey, R. Kotter, M. Breakspear, and O. Sporns, Proceedings of the National Academy of Sciences 104, 10240 (2007). [105] A. Neumaier and T. Schneider, ACM Trans. Math. Softw. 27, 27 (2001). [106] V. Jirsa and H. Haken, Physical Review Letters, Tech. Rep. 5 (1996). Neuroscience, Tech. Rep. 10 (2001). [107] J. G. T. Zanudo, G. Yang, and R. Albert, (2016), [94] S. Palva and J. M. Palva, Trends in Neurosciences 30, 10.1073/pnas.1617387114, arXiv:1605.08415. 150 (2007). [95] A. K. Engel and P. Fries, Current Opinion in Neurobi- ology 20, 156 (2010). [108] S. L. Brunton, J. L. Proctor, and J. N. Kutz, Pro- ceedings of the National Academy of Sciences 113, 3932 (2016).
1605.01270
1
1605
2016-05-04T13:26:30
Towards tailoring non-invasive brain stimulation using real-time fMRI and Bayesian optimization
[ "q-bio.NC" ]
Non-invasive brain stimulation, such as transcranial alternating current stimulation (tACS) provides a powerful tool to directly modulate brain oscillations that mediate complex cognitive processes. While the body of evidence about the effect of tACS on behavioral and cognitive performance is constantly growing, those studies fail to address the importance of subject- specific stimulation protocols. With this study here, we set the foundation to combine tACS with a recently presented framework that utilizes real-time fRMI and Bayesian optimization in order to identify the most optimal tACS protocol for a given individual. While Bayesian optimization is particularly relevant to such a scenario, its success depends on two fundamental choices: the choice of covariance kernel for the Gaussian process prior as well as the choice of acquisition function that guides the search. Using empirical (functional neuroimaging) as well as simulation data, we identified the squared exponential kernel and the upper confidence bound acquisition function to work best for our problem. These results will be used to inform our upcoming real- time experiments.
q-bio.NC
q-bio
Towards tailoring non-invasive brain stimulation using real-time fMRI and Bayesian optimization Romy Lorenz∗†, Ricardo P Monti∗†, Adam Hampshire†, Yury Koush‡, Christoforos Anagnostopoulos†, Aldo A Faisal†, David Sharp†, Giovanni Montana†§, Robert Leech† and Ines R Violante† †Imperial College London, UK, ‡EPFL, Switzerland, §King's College London, UK ∗ These authors contributed equally to this work. Abstract-Non-invasive brain stimulation, such as transcranial alternating current stimulation (tACS) provides a powerful tool to directly modulate brain oscillations that mediate complex cognitive processes. While the body of evidence about the effect of tACS on behavioral and cognitive performance is constantly growing, those studies fail to address the importance of subject- specific stimulation protocols. With this study here, we set the foundation to combine tACS with a recently presented framework that utilizes real-time fRMI and Bayesian optimization in order to identify the most optimal tACS protocol for a given individual. While Bayesian optimization is particularly relevant to such a scenario, its success depends on two fundamental choices: the choice of covariance kernel for the Gaussian process prior as well as the choice of acquisition function that guides the search. Using empirical (functional neuroimaging) as well as simulation data, we identified the squared exponential kernel and the upper confidence bound acquisition function to work best for our problem. These results will be used to inform our upcoming real- time experiments. Keywords-Bayesian optimization, fMRI, non-invasive brain stimulation, transcranial alternating current stimulation I. INTRODUCTION Studies involving non-invasive brain stimulation have re- ported remarkable changes in cognitive and behavioral perfor- mance [1]. Among those techniques, transcranial alternating current stimulation (tACS) is particularly promising as it directly allows for the modulation of physiologically relevant brain oscillations that subserve cognitive operations [2]. The effects of tACS are highly dependent on the frequency and phase of stimulation [3]. The conventional tACS approach involves defining the frequency and phase of stimulation ad hoc and testing them on a cohort of subjects. However, this approach exhibits two limitations: (1) the brain networks targeted by the stimulation cannot be verified without si- multaneous fMRI; (2) those stimulation parameters may vary across subjects due to difference in anatomy or heterogeneity in disease. Yet, there is a combinatorial explosion in the biologically plausible range of stimulation frequencies (1-100 Hz) and phases (0-359◦), resulting in up to thousands of possibilities. Identifying the optimal stimulation protocol for a given individual is like "finding a needle in a haystack". There- fore, using conventional methodology makes tailoring tACS to an individual highly unfeasible. To address this fundamental challenge in the clinical use of tACS, we aim to combine tACS with a recently presented framework that behavior real-time functional magnetic resonance imaging (fMRI) and Bayesian optimization: The Automatic Neuroscientist [4]. Employing Fig. 1. High-level overview of the Automatic Neuroscientist combined with tACS. (1) The experiment starts with applying a random combination of tACS parameters to the subject in the scanner. (2) Whole-brain functional images are acquired and analyzed in real-time in response to the block of stimulation. (3) Information about the current brain state is extracted and (4) compared to the pre-defined target brain state. This result is then fed into the Bayesian optimization approach. (5) Based on this, the algorithm chooses a new combination of tACS parameters that optimizes for the target brain state. This closed-loop cycle continues until the optimal tACS parameters are found. such a framework allows us to start with a target brain state and find a set of tACS parameters (frequency and phase of stimulation) for each individual that maximally activates it (Fig. 1). Bayesian optimization is particularly relevant to such a problem as it provides a powerful strategy for finding maxima of objective functions that are expensive to evaluate and might contain noisy measurements (both criteria are met with func- tional neuroimaging data). Moreover, Bayesian optimization is suited in situations where we do not have an analytical expres- sion of the objective function nor can make formal statements regarding its properties [5], [6]. However, the implementation of Bayesian optimization requires several fundamental choices that are paramount to the success of the technique; such as (1) the choice of covariance function for the Gaussian process (GP) prior and (2) the choice of acquisition function which determines the manner in which the space of parameters is explored. The aim of this study is to objectively compare the per- formance of a variety of distinct kernel functions (resulting in distinct GP priors) and acquisitions functions. Offline fMRI data from eight healthy volunteers receiving blocks of non- invasive tACS stimulation over left frontoparietal brain regions as well as simulation data were employed. The results obtained will be used to propose a combination of kernel and acquisition functions for upcoming real-time experiments. II. METHODS The objective of this work is to obtain a better under- standing of the relationship between various combinations of tACS parameters (i.e., frequency and phase) and each subject's neural response. In the framework of Bayesian optimization, such a relationship is summarized by a latent objective function which we wish to infer. With this in mind, a small neuroimag- ing study was conducted where healthy participants received tACS consisting of various combinations of frequency and phase. This data was subsequently employed to perform model selection on the GP covariance function. In addition to this, simulation analyses were carried out to compare two popular acquisition functions. A. Empirical Data 1) Subjects and Experimental Design: Eight healthy volun- teers (5 female, 24.75 ± 3.49 years) took part in our study. The study was approved by the Hammersmith Hospital (London, UK) research ethics committee. Each participant performed three runs of a switch task [7] in the MR scanner. Each run consisted of task blocks (36 s) interleaved with rest blocks (24 s). During task blocks, tACS was applied via a pair of MR-compatible conductive rubber electrodes over left frontal (F3) and left parietal (P3) regions (as determined by the International 10-20 EEG system). Both return electrodes were placed on the ipsilateral shoulder. The current was fixed to 1 mA (peak-to peak). The experiment parameter space was limited to ten different frequencies (0,1,5,8,12,16,20,26,40 and 80 Hz) and five different phases (0, 60, 90, 180 and 270◦), resulting in 50 different frequency-and-phase combinations (see Fig. 2). For each run and each subject, we tested a subset of 14 identical frequency-and-phase combinations (grey shaded squares in Fig. 2). The order of those frequency-and-phase combinations was random for each run in order to eliminate any order effect of the stimulation on the subject's neural response. The condition without any stimulation (0 Hz with 0◦) was the only condition that was repeated two (n=4) to four (n=4) times within each run. 2) Target brain state extraction: Whole-brain coverage images were acquired by a Siemens Verio 3T scanner using an EPI sequence (TR: 1.5s). Data was minimally preprocessed by applying motion-correction, high-pass filtering (100s) and spatial smoothing (5mm FWHM Gaussian kernel). We than ex- tracted 20 regions of interest (ROIs) that have been previously shown to predict trial-by-trial performance in task switching (for ROI definition we created 8mm spheres around the peak coordinates reported in [8]). The extracted timecourses were further cleaned by removing high-frequency noise and large signal spikes using a modified Kalman filter [9]. We then assessed changes in functional connectivity between those regions for blocks of tACS using psychophysiological inter- action (PPI) analyses [10] (whilst also including six motion parameters as confounds). For the analyses presented here, Exhaustive two-dimensional Fig. 2. tACS parameter space with each dimension corresponding to the frequency or phase of non-invasive brain stimulation, respectively. Grey shaded squares are combinations sampled in our (offline) fMRI study. we selected the two regions that exhibited the highest inter- run robustness across all subjects (assessed using Spearman's rank correlation coefficient): the left inferior parietal lobule and the posterior cingulate gyrus. In order to assess how likely the pattern of results between those regions could have occurred by chance, we performed non-parametric permutation testing (10,000 permutations) and found the result to be highly significant (z=3.71, p<.001). B. Bayesian optimization The underlying intuition behind the method of Lorenz et al. [4] is that the target brain pattern is a function of experimental conditions. As such, the authors propose to learn the relation- ship by modelling the observed brain state as a sample from a Gaussian process (GP). This facilitates the use of a Bayesian optimization framework [5], [6] in a closed-loop form where subjects are presented with an experimental condition and real-time fMRI provides instantaneous information about the subject's brain state. Based on this, we can iteratively update our beliefs about the unknown objective function, captured in the posterior distribution of the GP. While Lorenz et al. [4] used simple perceptual stimuli in their original experiment, our study here involves blocks of non-invasive brain stimulation with different combinations of frequency and phase. This can be considered as a far more challenging problem as only a handful of studies exist investigating the effects of tACS on the blood oxygenation level dependent (BOLD). Beyond that, high inter-subject variability can be expected as outlined in the Introduction. In contrast to the fMRI setting employed in [4], there is far less prior knowledge regarding the properties of the latent objective function which can be leveraged in this context. Therefore, data was collected over a cohort of eight subjects (see previous section). Using this empirical data we begin by studying the choice of kernel function in Section II-B1. Given a choice of kernel and by carrying out simulation analyses, the effects of various acquisition functions are subsequently assessed in Section II-B2. 1) Covariance function: The choice of kernel function in GP regression is fundamental [11]. The kernel directly spec- ifies a measure of similarity across various inputs and there- fore determines the generalization properties of the estimated model. Model selection in the context of GPs is often posed as a parameter estimation task whereby the hyper-parameters for a specific choice of kernel can be learnt [11]. However, comparing the performance of potentially many distinct kernels is a challenging problem. Recently, [12], [13] have proposed a method for automatically searching over the space of kernel structures in a principled manner. Briefly, the proposed approach proceeds by considering compositional structures (either addition or multiplication) over four base kernels; squared exponential (as employed in [4]), periodic, linear or rational quadratic. When comparing various distinct kernel structures the Bayesian Information Cri- terion (BIC) is employed to score the various models (where type-2 maximum likelihood estimates of parameters are used). Initially, each of the base kernels is proposed. Thereafter, the selected kernel can be extended by either performing addition or multiplication with any base kernel. While the original approach described in [12] proposes kernels for each dimension independently, in this work both dimensions were studied simultaneously (thus two dimensional kernels were proposed at each step). This decision is based on the hypothesis that each dimension (either frequency or phase) would share a similar relationship with the response, therefore leading to the selection of similar (if not the same) kernel. 2) Acquisition function: Of equal importance for the suc- cess of the Bayesian optimization is the choice of acquisition function. The role of the acquisition function is to iteratively propose the experimental condition to be presented to the subject. It therefore serves to guide the search over all com- binations of parameters (in our case frequency and phase) and must implicitly balance exploration with exploitation. In this work we considered the performance of two popular acquisition functions: the expected improvement (EI) and upper confidence bound (GP-UCB) acquisition function [5]1. The two acquisition functions differ in their trade-off between exploration and exploitation of the experimental parameters space, giving rise to distinct sampling behaviors over time. Informally, the EI acquisition function can be seen as trying to maximize the expected improvement over the current best. While the GP-UCB also favors the selection of points with high mean value (similar to the EI acquisition function), it also favors points with high variance (i.e., regions worth to explore). In certain settings, in a more explorative behavior of the GP-UCB compared to the more "greedy" EI acquisition function. For algorithmic details, please refer to [5]. this will result While the choice of kernel function can be guided by model selection and information theoretic measures, the comparison of multiple acquisition functions requires explicit knowledge of the underlying objective function. As a result, a simulation study was employed to the benchmark the performance of the two aforementioned acquisition functions. A complex and multimodal objective function was pro- posed for the simulation study (see Fig. 3a). This was moti- vated by our empirical results, showing multiple optima and minima for each subject (not shown). As the presence of non-neural noise is well documented for fMRI experiments, we also studied how different levels of contrast-to-noise-ratio 1 The probability of improvement (PI) acquisition function was not included as it is typically seen to be over exploitative. (CNR) affected the Bayesian optimization. We repeated our simulations analyses with CNR values ranging from 0.1 and 1.8 (typically reported CNR values in fMRI literature vary between 0.5 and 1.8 [14]). For each CNR value tested, we ran 100 simulations. The maximum number of iterations was set to 100 with the first five (randomly selected) observations serving as burn-in phase. Thereafter, at each iteration, a new frequency-and-phase combination was proposed by maximiz- ing the respective acquisition function and sampled by the Bayesian optimization algorithm in the next iteration. As a measure of accuracy we computed spatial correlation between the algorithm's predictions for the whole parameter space and the "ground truth" objective function. Considering our underlying motivation to gain a holistic understanding of the whole parameter space (learning about maxima and min- ima), spatial correlations seemed most appropriate to capture similarity across the whole space. This procedure was identical to the one reported in [4]. III. RESULTS A. Covariance kernel selection In this work we follow [12], [13], and look to compare several distinct kernels via the use of BIC. Furthermore, we also employ a greedy search algorithm whereby we begin with a base kernel and iteratively extend this kernel. linear and Matern ( 3 There are two significant differences in the approaches described in this work to those of [12], [13]. The first is our choice of base kernels; in this work the squared exponential, periodic, 2 ) where employed as base kernels. The second difference is that two-dimensional kernels were proposed at each iteration. This choice was motivated by an assumption that the relationship between parameter (fre- quency or phase) and the response shared a similar functional form. Finally, a minor difference between this work and that of [13] is that change-point functions were not considered here. Following [12], we begin by proposing each of the base kernels and calculating the associated BIC scores. In this first step the squared exponential kernel was selected. We then proceeded to consider compositional structures. Each of these structures was composed by either adding or multiplying the squared exponential kernel with each of the base kernels. For each compositional kernel, the parameters were estimated using the previous parameters as warm starts. Again, BIC scores were employed to score each of the compositional models. At this stage, there was no reduction in BIC scores resulting in the choice of the squared exponential kernel. B. Acquisition function selection Simulation analyses were carried out to assess the perfor- mance of the EI and GP-UCB acquisition functions. A range of CRN values were employed in order to recreate the properties of fMRI data as well as study the performance of each of the acquisition functions in scenarios with low signal-to-noise ratio. Throughout this simulation the underlying GP model was maintained constant, allowing us to attribute any differences in performance to the acquisition functions. For each CNR value, 100 simulations where performed. Each simulation allowed the proposed Bayesian optimization Fig. 3. Results of simulation analyses. (a) Modeled objective function used for simulations. Mean ± SEM (shaded areas) spatial correlation between predicted and modeled parameter space for (b) EI and (c) GP-UCB acquisition function. As the first five iterations were used as a burn-in, they are not depicted here (gray dashed line). Simulations were performed for 10 different CNR values, ranging from 0.1 (bright blue) to 1.8 (dark blue). algorithm to explore the parameter space over 100 iterations. The spatial correlation between the predictive posterior under the GP model and the true objective function was calculated at each iteration. The mean ± SEM spatial correlation results are depicted in Fig. 3b for the EI and in Fig. 3c for the GP-UCB acquisition function. Not surprisingly, we note that both acquisition func- tions fail in the presence of too much noise (for CNR values < 0.2). The results also indicate that the GP-UCB acquisition function clearly outperforms the EI acquisition function for typical CNR values reported in the literature (0.5-1.8). IV. DISCUSSION AND FUTURE WORK The results presented here serve as an initial exploration into the choices of kernel and acquisition functions which can be used in the context of real-time optimization for tACS. Our results propose the use of a squared exponential kernel function in combination with a GP-UCB acquisition function. In a series of simulations, the latter was shown to capture the whole parameter space (as identified by using spatial correlations) when compared to the EI acquisition function. This result was consistent for a range of CNR values tested. These findings serve as an important basis for upcoming real- time experiments. An avenue of future work is the definition of appropriate stopping criteria to determine convergence [15]. This is particularly important in our context due to high scanning costs, limited attentional capacities of subjects and neural habituation to the stimulation. To our knowledge, this is the first study to propose a framework that allows tailoring non-invasive brain stimulation to an individual. The next step consists of applying this technology to patients suffering from traumatic brain injury; a condition that disrupts frontoparietal brain networks and hence results in cognitive impairments [16]. While closed-loop deep brain stimulation has already been shown to outperform the conventional approach in Parkinson's patients [17] we envision that our framework will advance personalized treatment by means of non-invasive brain stimulation. This will be of particular importance for neurological and psychiatric deficits that are diffuse and widely heterogeneous in their origin. ACKNOWLEDGMENT This research was supported by the NIHR Imperial BRC and the Wellcome Trust (103045/Z/13/Z). REFERENCES [1] MF Kuo, MA Nitsche: Effects of transcranial electrical stimulation on cognition, Clin. EEG Neurosci.,43(3), 92–199 (2012) [2] CS Herrmann, S Rach, T Neuling, D Strueber: Transcranial alternating current stimulation: a review of the underlying mechanisms and modulation of cognitive processes, Frontiers Human Neuroscience, 7:279 (2013) [3] A Antal, W Paulus: Transcranial alternating current stimulation (tACS), Frontiers Human Neuroscience, 7:317 (2013) [4] R Lorenz, RP Monti, IR Violante et al.: The Automatic Neuroscientist: A framework for optimizing experimental design with closed-loop real-time fMRI. NeuroImage,129, 320–334 (2016) [5] E Brochu, VM Cora, N de Freitas: A tutorial on Bayesian optimization of expensive cost functions, with application to active user modeling and hierarchical reinforcement learning. arXiv:1012.2599 (2010) [6] B Shahriari, K Swersky, Z Wang, RP Adams, N de Freitas: Taking the Human Out of the Loop: A Review of Bayesian Optimization, Proceedings of the IEEE, 104(1), 148–175 (2016) [7] A Dove, S Pollmann, T Schubert, CJ Wiggins, DY von Cramon: Prefrontal cortex activation in task switching: an event-related fMRI study. Cognitive Brain Research, 9(1): 103–109 (2000) [8] AB Leber, NB Turk-Browne, MM Chun: Neural predictors of moment-to-moment fluctuations in cognitive flexibility, PNAS, 105(36), 13592–13597 (2008) [9] Y Koush, M Zvyagintsev, M Dyck, KA Mathiak, K Mathiak: Signal quality and Bayesian signal processing in neurofeedback based on real-time fMRI, NeuroImage, 59(1) (2012) [10] KJ Friston: Functional and Effective Connectivity: A Review, Brain Connectivity, 1(1),13–36 (2011) [11] CE Rasmussen, CK Williams: Gaussian processes for machine learning. MIT Press (2006) [12] D Duvenaud, JR Lloyd, R Grosse, JB Tenenbaum, Z Ghahramani.: Structure discovery in nonparametric regression through compositional kernel search, ICML (2013) JT Lloyd, D Duvenaud, R Grosse, JB Tenenbaum, Z Ghahramani.: Automatic construction and natural-language description of nonparametric regression models, Association for the Advancement of Artificial Intelligence (2014) [13] [14] M Welvaert, Y Rosseel: On the definition of signal-to-noise ratio and contrast-to- noise ratio for fMRI data, PLoS One, 8, doi: 10.1371/journal.pone.0077089 (2013) [15] R Lorenz, RP Monti, IR Violante et al.: Stopping criteria for boosting au- tomatic experimental design using real-time fMRI with Bayesian optimization, arXiv:1511.07827 (2015) SR Jilka, G Scott, T Ham, A Pickering, V Bonnelle, RM Braga, R Leech, DJ Sharp: Damage to the Salience Network and Interactions with the Default Mode Network, J. Neurosci., 34(33), 0798-10807 (2014) S Little, A Pogosyan, S Neal et al.:Adaptive deep brain stimulation in advanced Parkinson disease, Ann. Neurol., 74(3), 449–457 (2013) [16] [17]
1503.07610
1
1503
2015-03-26T03:44:43
Eye-Movement Control During the Reading of Chinese: An Analysis Using the Landolt-C Paradigm
[ "q-bio.NC" ]
Participants in an eye-movement experiment performed a modified version of the Landolt-C paradigm (Williams & Pollatsek, 2007) in which they searched for target squares embedded in linear arrays of spatially contiguous "words" (i.e., short sequences of squares having missing segments of variable size and orientation). Although the distributions of single- and first-of-multiple fixation locations replicated previous patterns suggesting saccade targeting (e.g., Yan, Kliegl, Richter, Nuthmann, & Shu, 2010), the distribution of all forward fixation locations was uniform, suggesting the absence of specific saccade targets. Furthermore, properties of the "words" (e.g., gap size) also influenced fixation durations and forward saccade length, suggesting that on-going processing affects decisions about when and where (i.e., how far) to move the eyes. The theoretical implications of these results for existing and future accounts of eye-movement control are discussed.
q-bio.NC
q-bio
Eye-Movement Control During the Reading of Chinese: An Analysis Using the Landolt-C Paradigm Yanping Liu1, Erik D. Reichle2, & Ren Huang3 1Key Laboratory of Behavioral Science Institute of Psychology, Chinese Academy of Sciences 2School of Psychology, University of Southampton, UK 3Department of Psychology, Sun Yat-Sen University, China Author note This research was supported by two grants from the China Postdoctoral Science Foundation (2013M541073 & 2014T70132) awarded to the first author, and by a Grant HD075800 from the National Institutes of Health awarded to the second author. Correspondence should be addressed to Yanping Liu, 16 Lincui Road, Key Laboratory of Behavioral Science, Institute of Psychology, Chinese Academy of Sciences, Beijing, China. Email: [email protected]. 1 Abstract Participants in an eye-movement experiment performed a modified version of the Landolt-C paradigm (Williams & Pollatsek, 2007) in which they searched for target squares embedded in linear arrays of spatially contiguous “words” (i.e., short sequences of squares having missing segments of variable size and orientation). Although the distributions of single- and first-of-multiple fixation locations replicated previous patterns suggesting saccade targeting (e.g., Yan, Kliegl, Richter, Nuthmann, & Shu, 2010), the distribution of all forward fixation locations was uniform, suggesting the absence of specific saccade targets. Furthermore, properties of the “words” (e.g., gap size) also influenced fixation durations and forward saccade length, suggesting that on-going processing affects decisions about when and where (i.e., how far) to move the eyes. The theoretical implications of these results for existing and future accounts of eye-movement control are discussed. Keywords: Chinese reading; eye-movement control; Landolt-C paradigm; lexical processing; reading; saccade targeting 2 Neural systems are extremely adept at exploiting the regularities that exist in the environment for the purposes of making (near) optimal inferences to guide behavior (Anderson, 1990). For example, the systems that mediate vision exploit regularities of objects (e.g., the Gestalt principle of similarity) to represent those objects in a manner that affords their invariant perception despite variability of both the objects and viewing angle, partial occlusion of the objects, etc. (Smith, 1988). It should therefore come as little surprise that complex visual tasks like reading also exploit regularity, allowing the task to be performed in a (near) optimal manner (Liu & Reichle, 2010; Liu, Reichle, & Gao, 2013; Reichle & Laurent, 2006; see also Bicknell & Levy, 2012). For example, because low-frequency words require more time to identify than high-frequency words (Forster & Chambers, 1973; Schilling, Rayner, & Chumbley, 1998), readers of alphabetic languages like English tend to spend more time fixating the low- than high-frequency words (Inhoff & Rayner, 1986; Kliegl, Nuthmann, & Engbert, 2006; Rayner, Ashby, Pollatsek, & Reichle, 2004). And in a similar manner, because words can be identified most rapidly when they are fixated near their center (O’Regan, 1981; Rayner, 1979; Rayner & Morrison, 1981), readers of alphabetic languages use the blank spaces between words to direct their eyes to locations near the center of words (McConkie, Kerr, Reddix, & Zola, 1988; Rayner, Sereno, & Raney, 1996). Such eye-movement behavior takes advantage of the inherent regularities of the text to support reading that is rapid, but that also maintains some overall level of comprehension. What is less clear, however, is whether such behaviors are universal, generalizing across languages and writing systems that exhibit less—or perhaps different—patterns of regularity. One prime example that has been the focus of much recent research is Chinese (for a review, see Zang, Liversedge, Bai, & Yan, 2011). 3 The Chinese language and writing system has many properties that make it different from the alphabetical languages that have most often been used in experiments to understand reading. For example, in contrast to alphabetic writing systems, Chinese words are not comprised of letters and are not spatially segregated by blank spaces. Instead, Chinese words consist of 1-4 equally sized box-shaped characters that are comprised of 1-36 “strokes” (i.e., line segments that originally corresponding to the brush strokes used to write the characters), with the characters being arranged in a spatially adjacent linear array (see Figure 1 for an example). The fact that Chinese words are comprised of a variable number of characters and not demarcated by blank spaces means that Chinese readers must use their knowledge of their language to somehow segment character sequences into their corresponding words (e.g., see Li, Rayner, & Cave, 2009). This task of segmenting characters into words is intrinsically difficult because there is often ambiguity in how a character sequences might be segmented, with a given sequence of four characters, for example, possibly corresponding to two 2-character words or one 4-character word (Hoosain, 1991, 1992; Liu, Li, Lin, & Li, 2013). ----- Figure 1 ----- Because of these properties of written Chinese words, there is some debate about both the extent to which and how the eye movements of Chinese readers are influenced by the lexical properties of words (as they clearly are in alphabetic languages; see Rayner, 1998, 2009). For example, one theoretical “camp” has argued that characters are more important than words during the reading of Chinese, and that characters are the basic units that determine both when the eyes move (Chen, 1996; Chen, Song, Lau, Wong, & Tang, 2003; Chen & Zhou, 1999; Feng, 2008; Hoosain, 1991, 1992) and where the eyes move (Tsai & McConkie, 2003; Yang & McConkie, 4 1999). In contrast, the other theoretical “camp” maintains that, because words have a psychological reality in the minds of Chinese readers (Cheng, 1981; Li et al., 2009; Li, Bicknell, Liu, Wei, & Rayner, 2014; Li & Logan, 2008; Li, Gu, Liu, & Rayner, 2013), they also play an important functional role in determining when the eyes move during the reading of Chinese (Liversedge, Hyönä, & Rayner, 2013; Yan, Tian, Bai, & Rayner, 2006). However, this second perspective further diverges in their accounts of how readers of Chinese decide where to move their eyes. One contingent has argued that the locations of initial fixations on words are a function of whether the words are successfully segmented in the parafovea, with saccades towards segmented words being directed towards their centers but saccades towards un-segmented words being directed towards their beginnings (Yan et al., 2010). In contract, the other contingent has proposed that there is no “default” saccade target within a word, but that the length of the saccade moving the eyes into a word is instead influenced by the relative difficulty of foveal and/or parafoveal processing, with easier processing resulting in longer forward saccades (Liu, Reichle, & Li, 2014; Wei, Li, & Pollatsek, 2013). Given these different theoretical perspectives on eye-movement control during the reading of Chinese, it is clear that—in contrast to alphabetic languages—we still know very little about the basic mechanisms that determine when and where the eyes move during the reading of non-alphabetic languages, or more specifically, about if and, if so, how lexical processing mediates eye-movement control during the reading of Chinese. One strategy that has been used to examine the influence of lexical processing in guiding eye movements during the reading of alphabetic languages has involved comparisons of such eye movements to those that are observed during tasks that do not require lexical processing but that do entail many of the same visual and oculomotor constraints, such as the Landolt-C “reading” task (Corbic, Glover, & 5 Radach, 2007; Williams & Pollatsek, 2007; Vanyukov, Warren, Wheeler, & Reichle, 2012; see also Williams, Pollatsek, & Reichle, 2014). In experiments that have used this task, participants were instructed to scan through linear arrays of Landolt Cs (i.e., ring-shaped stimuli having missing segments (i.e., gaps) of variable size and/or orientations) to search for target stimuli—rings containing no missing gaps. Because these linear arrays are comprised of Landolt-C stimuli arranged into sequences of “words” (i.e., clusters of the stimuli separated by blank spaces between clusters), the task has been used to study saccadic targeting under conditions that resemble those experienced during the reading of alphabetic languages, minus all of the language- related behaviors that are normally engaged during word, sentence, and discourse processing. The key findings from these experiments is that participants seem to treat the Landolt-C clusters as “words”, directing their eyes towards their centers and fixating the easier-to-process clusters (i.e., those with larger gaps sizes and/or that occur more often during the course of the experiment) for shorter durations than the more difficult-to-process clusters (i.e., those with smaller gap sizes and/or that occur less often during the experiment). In the experiment that is reported in this article, the Landolt-C task was modified so that it more closely resembled the reading of Chinese (see Figure 2), thereby providing a novel way to examine if the segmentation of “words” (i.e., spatially adjacent stimuli having similarly sized and orientated gaps) affects fixation durations and/or saccadic targeting1. That is, our participants were instructed to scan linear arrays of Landolt-squares, or boxes with a missing segment of varying size and orientation, and to indicate the number of targets (i.e., squares that do not contain missing segments; e.g., □). The advantages of this method are that it provides a way of precisely manipulating both word processing difficulty (i.e., gap size) and size (i.e., 6 the number of identical spatially adjacent non-target squares) in the absence of language processing. The main goal of this experiment was therefore to determine if/how word segmentation and identification affects eye movements in this task so as to better inform our understanding of the word-based effects that have been observed during the actual reading of Chinese text. More specifically, we focus on: (1) if/how fixation durations on individual words are affected by their identification difficulty; and (2) if/how saccadic targeting is affected by their segmentation difficulty. If easier-to- identify words are fixated for shorter durations but saccade targeting is unaffected by the ease of word segmentation, then this pattern of results would lend further support to the hypothesis that the processing associated with words influences both when and where the eyes move during the reading of Chinese (e.g., Liu et al., 2014; Wei et al., 2013), but that parafoveal word segmentation is not the primary determinant of where the eyes are moved (cf., Yan et al., 2010). Method Participants Twenty participants (12 males) recruited from Sun Yat-Sen University were paid for their participation. All participants had normal or corrected-to-normal vision and were naive to the purpose of the experiment. Apparatus Stimuli were displayed on a 23-in. LCD monitor (Samsung SyncMaster2233) using SR-Research Experiment Builder software. The monitor had a resolution of 1,680 × 1,050 pixels and refresh rate of 120 Hz. Eye movements were recorded using 7 a SR-Research Eyelink 1000 eye tracker (Kanata, ON, Canada) with a spatial resolution of 0.01° and sampling rate of 1,000 Hz. Stimuli and Design The materials were designed to be as similar to Chinese sentences as possible (e.g., cf., Figures 1 vs. 2). Each “character” was a Landolt-Square (40 pixels × 40 pixels) with a gap that was 2, 4, 6, or 8 pixels in size and that occurred in the left, right, top, or bottom of the character. Each cluster or “word” was composed of 1, 2, 3, or 4 characters. Both the gap size and location were held constant within a given word, but varied between words. The spaces between any two successive characters were 6 pixels. Each array of words or “sentence” contained 10 randomly selected words and ranged from 16 to 33 characters in length (M = 25 characters). The combination of the number of characters per word and the gap size and orientation in a given word gave 64 unique words. Each word was repeated 40 times across the experiment. Each sentence had 0, 1, or 2 targets (i.e., characters without gaps) that could appear with equal probability within any word except the first and last word within a sentence. Participants were instructed to scan through each sentence and then indicate via button presses how many targets occurred in the sentence, with one sentence being displayed per trial. There were four blocks containing 64 trials each, with the order of blocks being completely random. ----- Figure 2 ----- Procedure Upon arrival, participants were given task instructions and then seated 63 cm from the video monitor. A chinrest was used to minimize head movements during the experiment. Although viewing was binocular, eye-movement data were only collected 8 from right eye. The eye-tracker was calibrated and validated using a 9-dot procedure at the beginning of each block of trials, with additional calibrations and validations being conducted as necessary. After that, participants completed 8 practice trials (not included in our analyses) and then completed the 4 blocks of experimental trials. A drift-check procedure was performed before each trial; a sentence was displayed after the participants successfully fixated the white box (1° × 1°) at the location of the first character in the sentence. Participants were instructed to view a right-bottom region of the screen to terminate a trial and to then press the number button on the keyboard corresponding to the number of targets that appeared in the trial. Accuracy feedback was randomly provided on half of the trials. Results Accuracy The mean overall accuracy of the experiment was 91% (SD = 0.05). Eye-Movement Results Fixations on the first and last words in each sentence were removed from our analyses because the former coincided with the abrupt appearance of the sentence and the latter coincided with the termination of a trial. Fixations on words containing targets and immediately preceding or following targets were also removed from our analyses. Thus, fixations on 64.3% of the total number of words were included in our analyses. From these data, saccades longer than 3 SDs above the mean for a given participant were excluded (3.05% of the total saccades). To determine if/how the processing of words affected eye movements during our task, our analyses focused on the relationship between various eye-movement dependent measures (related to when and where to move eyes) and properties of both 9 the fixated word (i.e., word N) and the two spatially adjacent words (i.e., words N-1 and N+1). To do this, we examined the first-pass fixation-location distributions corresponding to: (1) first fixations; (2) single fixations; (3) first-of-multiple fixations; and (4) all forward fixations. We also examined the following measures: (5) forward- saccade length, or the distance between two consecutive forward fixations during first-pass scanning; (6) first-fixation duration, or the duration of the first fixation on a word during first-pass scanning; (7) gaze duration, or the sum of all first-pass fixation durations on a word; and (8) total-viewing time, or the sum of all fixation durations on a word. We then built linear mixed-effects models (LMMs) for each measure, specifying participants, items, and random slopes of each predictor (e.g., word properties) as the random effects so that our reported significance values reflect the variability of participants, items, and the slopes of the predictors (see also Barr, Levy, Scheepers, & Tily, 2013). Based on prior results and our priori hypotheses, we also included practice (i.e., ordinal trial number) and the properties of words N-1, N, and N+1 (i.e., gap size and number of characters) as fixed-effect factors in our LMMs. The models were then fitted using the lme4 package (ver. 1.1-7; Bates, Maechler, Bolker, & Walker, 2014; Pinheiro & Bates, 2000) and p-values were estimated using lmerTest package (ver. 2.0-20; Kuznetsova, Brockhoff, & Christensen, 2013) in R (ver. 3.1.2; R Development Core Team, 2015). Fixation Locations To determine if parafoveal word segmentation influences saccadic targeting, we analyzed our various measures of fixation location during first-pass scanning. Figure 3 respectively shows the distributions of initial-, single-, first-of-multiple-, and all forward fixation locations as a function of the number of characters with the words, their gap sizes, and practice (i.e., ordinal block number). As shown, the locations of 10 initial fixations (panels a-b) and the first-of-multiple fixations (panels c-d) were more likely to be near the beginnings of words, whereas single fixations (panels e-f) were more likely to be located near the centers of words. However, the locations of all forward fixations (panels g-h) were uniform (all values of χ2 ≤ 1.74, ps ≥ 0.755; also see Table 1). All of these results are consistent with previous findings that have been observed with Chinese reading (Li, Liu, & Rayner, 2011; Yan et al., 2010). ----- Figure 3 & Table 1 ----- LMMs were further used to examine if practice and the properties of words N-1, N, and N+1 affected all forward fixation locations on word N (measured from the left edge of that word). Unsurprisingly, Table 2 shows that the mean fixation location across all forward saccades moved further to the right of word N as its length increased (b = 0.99, SE = 0.01, t = 97.27, p < 0.001). The fact that the mean fixation location was not influenced by the gap size of word N, the properties of the spatially adjacent words, or the amount of practice (all values of t ≤ 1.24, all ps ≥ 0.214) indicates that the uniformity of distributions of forward fixation locations was robust, which in turn suggests the absence of any (strongly) preferred viewing location on the words in our task. Importantly, the “preferred viewing locations” which were seemingly evident in the distributions of first and first-of-multiple fixations (i.e., located near the word beginnings) and of the single fixations (i.e., located near the word centers) were specious because of the arbitrary way in which these fixations were extracted from the uniform distribution of all forward fixations. We will say more about this in the Discussion. ----- Table 2 ----- Saccade Length 11 Although the results from fixation-location analyses indicated that properties of neither the fixated word nor its spatially adjacent words influenced the distribution of all of the forward fixation locations, it is possible that foveal and/or parafoveal processing influenced the length of forward saccade. As Table 3 indicates, forward saccade length increased with the lengths of word N-1 (b = 0.01, SE = 0.01, t = 2.43, p = 0.015), word N (b = 0.10, SE = 0.01, t = 9.88, p < 0.001), and word N+1 (b = 0.04, SE = 0.01, t = 5.05, p < 0.001). Forward saccade length also increased with the gap size of word N-1 (b = 0.01, SE = 0.003, t = 1.72, p = 0.085), word N (b = 0.01, SE = 0.01, t = 2.23, p = 0.037), and word N+1 (b = 0.02, SE = 0.003, t = 5.58, p < 0.001). These results were entirely consistent with those of previously reported experiments involving Chinese reading which demonstrated that forward saccade length increases with both the processing ease (frequency) and length of the fixated word (Li et al., 2014; Liu et al., 2014; Liu, Li, &Pollatsek, submitted for review; Wei et al., 2013), and with the processing ease (frequency) and length of upcoming (parafoveal) word (Li et al., 2014; Liu et al., 2014; Liu et al., submitted for review). Fixation Durations ----- Table 3 ----- Our analyses have shown that properties of the fixated words and its spatially adjacent words did not influence the uniformity of all forward fixation location distributions, but that word properties did influence the length of forward saccades. This suggests that ongoing processing influences saccade length rather than fixation locations on words. It is therefore also necessary to examine how word properties and practice affected ongoing processing. As Table 4 shows, first-fixation durations decreased with the length of word N (b = -1.35, SE = 0.67, t = -2.01, p = 0.051) and the gap size of word N (b = -1.16, SE = 12 0.33, t = -3.50, p = 0.002). Conversely, as Table 5 shows, gaze durations increased with the length of word N (b = 135.70, SE = 10.91, t = 12.44, p < 0.001), but decreased with the lengths of word N-1 (b = -4.03, SE = 1.21, t = -3.34, p < 0.001) and word N+1 (b = -6.84, SE = 1.53, t = -4.46, p < 0.001). Similarly, gaze durations also decreased with increasing gap size of word N (b = -9.75, SE = 0.72, t = -13.47, p < 0.001). ----- Table 4 & 5 ----- Finally, as Table 6 shows, total-viewing times (like gaze durations) increased with the length of word N (b = 158.70, SE = 3.66, t = 43.37, p < 0.001), but decreased with the lengths of word N-1 (b = -5.06, SE = 1.41, t = -3.59, p < 0.001) and word N+1 (b = -6.53, SE = 1.41, t = -4.62, p < 0.001). And like first-fixation and gaze durations, the total-viewing times deceased with increasing gap size of word N (b = -13.16, SE = 0.85, t = -15.44, p < 0.001). ----- Table 6 ----- The results of the preceding analyses are consistent with other experimental results showing that the properties of the fixated word and its spatially adjacent neighbors can influence fixation durations during Chinese reading. For example, several studies have demonstrated that fixation durations decrease with increasing processing ease (e.g., frequency) of both the fixated word (Yan et al., 2006; Yang & McConkie, 1999; Rayner, Li, Juhasz, & Yan, 2005) and the previously fixated word (i.e., spillover effects; Li et al., 2014, see Table 1). Our results are also consistent with evidence that gaze durations tend to increase as the length of the fixated word increases and as the length of the previously fixated word decreases (Li et al., 2014, 13 see Table 1). The theoretical implications of these results for our understanding of eye-movement control during the reading of Chinese will be discussed next. General Discussion The present experiment aimed to determine if and how “word” processing affects eye movements in a Chinese-like visual search task so as to better inform our understanding of the word-related effects that have been observed in the actual reading of Chinese text. We replicated the well-established pattern of fixation locations that has been reported in Chinese reading: Whereas the distribution of first- fixation locations peaked at the word beginnings when the words were fixated more than once, the distribution of single-fixation locations peaked near the word centers (Li et al., 2011; Yan et al., 2010). However, an examination of the distributions of all of the forward fixation locations indicated that they were clearly uniform and not influenced by word length, word-processing difficulty (i.e., gap size), or practice. Finally, there were effects of word length and word-processing difficulty on the various fixation-duration measures and forward saccade length, suggesting that variables that influence that rate of on-going word processing also dynamically influences both when and where the eyes move during our task and—as we will argue, by extension—the reading of Chinese text (Liu et al., 2014). In the remainder of this discussion, we indicate precisely why we maintain this position, and discuss in more detail what we believe each of the preceding results says about how readers of Chinese make “decisions” about when and where to move their eyes. First, the finding that the distributions of all forward fixation locations were uniform and invariant to both “word” properties and practice strongly suggests that there is no preferred viewing location in our task; otherwise, the distributions would have tended to be normal, peaked on whatever viewing location actually afforded 14 efficient performance of our task. We believe that the patterns of fixation-location distributions observed for first-of-multiple and single fixations in our study and others (e.g., Li et al., 2011) and that have been suggested as indicative of a preferred viewing location that is conditional upon the successful segmentation of parafoveal words (e.g., see Yan et al., 2010) is instead an artifact of the simple fact that words that are fixated near their centers are less likely to be refixated than words that are fixated nears their beginnings (Liu et al., submitted for review). Li et al. (2011) provide convincing support for this conjecture by running simulations of eye movements during Chinese reading that included the simple assumption that saccades are of constant mean length but with some variability. These simulations reproduced the findings that the distributions of single- and first-of-multiple fixation locations tend to be near the centers and beginnings of words, respectively, thereby demonstrated precisely how a pattern of results that might otherwise suggest a type of saccade targeting that is conditional upon parafoveal word segmentation can emerge from simple assumptions that involve no specific saccade targets (see also Ma, Li, & Pollatsek, in press). These simulations, in conjunction with our findings, are therefore consistent with the hypothesis that there are no specific saccade targets during the reading of Chinese. In other words, a saccade target is not a true target per se, but is instead the consequence of a saccade that is intended to move the eyes to an informative viewing location. Second, the finding that properties of words influenced the various fixation- duration measures in our task is completely consistent with findings that lexical properties of Chinese words also influence such measures during the reading of Chinese (Li et al., 2014; Liu et al., 2014; Liu et al., submitted for review; Yan et al., 2006; Yang & McConkie, 1999; Rayner et al., 2005). These findings indicate the 15 validity of our method of using our version of the Landolt-C paradigm to examine eye-movement control during Chinese reading by replicating key results showing that the relative difficulty of “word” processing (Williams & Pollatsek, 2007; Williams et al., 2014) can influence the time spent fixating (processing) a given “word”, analogous to what is known to happen in actual reading (e.g., Schilling et al., 1998). These findings are therefore important because they permits one to infer that whatever perceptual (e.g., segmenting individual “words” based on character gap size and orientation) and cognitive (e.g., discriminating non-targets from targets) are operative in our task are likely to have analogs in Chinese reading (e.g., segmenting individual words, identifying characters and words, etc.), and that, in both tasks, these processes are likely to play causal roles in the patterns of eye movements that are observed. The most transparent of these causal roles is undoubtedly related to the moment-by- moment “decisions” about when to move the eyes from one location to the next, but as our experiment demonstrates, the causal influence is also likely to extend to decisions about where to move the eyes. This last conjecture is consistent with our third main finding that the properties of “words” influences forward saccade length in our task. On some level, this finding should not be too surprising if one considers the range on possible ways that readers might in theory guide their eyes while reading Chinese. On one end of the continuum, it is possible—thought not likely—that readers simply move their eyes randomly to new locations, perhaps adopting some simple heuristic (e.g., the “fixed saccade length” assumption used in the simulations reported by Li et al., 2011; also see Yan et al., 2010). At the other end of the continuum, readers might use word boundary information to move their eyes to the center of the next unidentified word, as posited to occur in models that simulate the eye movements of people reading alphabetic 16 languages like English and German (e.g., E-Z Reader: Reichle, Pollatsek, Fisher, & Rayner, 1998; Reichle, Warren, & McConnell, 2009; Reichle, 2011; SWIFT: Engbert, Nuthmann, Richter, & Kliegl, 2005; Schad & Engbert, 2012). This second possibility also seems unlikely given the empirical evidence against preferred viewing locations in both this experiment and in experiments involving Chinese reading (e.g., Li et al., 2011), and given the inherent difficulties associated with word segmentation in Chinese reading (e.g., see Li et al., 2009). By the process of elimination, then, the fact that these two extreme possible accounts of saccade targeting in Chinese are not very feasible suggests a third possibility—that readers decide where to move their eyes using some type of information other than word boundaries. What might this information be? One possibility that we believe is a viable candidate is whatever processing difficulty is associated with the currently fixated word and/or the word(s) immediately to the right of fixation. By this account, readers might use the foveal load, or the difficulty associated with identifying the fixated word, as a metric to gauge how far to move their eyes with the next saccade. The intuition behind this simple heuristic is that, because difficult-to-process fixated words afford less parafoveal preview of the upcoming word than do less difficult-to-process fixated words (Henderson & Ferriera, 1990; Kennison & Clifton, 1995), foveal load might provide the reader with a metric of how much processing still remains to be done on the parafoveal word when it is fixated, and thus a metric of how far the eyes should be moved into the word. (A parafoveal word that has received little processing during preview is likely to require a substantial amount of processing after it is fixated, making a more conservative fixation near the beginning of the word, thereby allowing for the possibility of one or more additional fixations on the word during the forward pass through the text.) And 17 for the same reason, information that becomes available about the parafoveal word from the foveal word might also be expected to play an important role in modulating the length of the saccade leaving the foveated word if the goal is to move the eyes just past the location where the parafoveal characters and/or words have been sufficiently processed for their identification to be imminent. Such a proposal has already been suggested to account for experimental results showing that the saccade length leaving a word is modulated by the difficulty (e.g., frequency) of that word during Chinese reading (Liu et al., 2014; Wei et al., 2013), and as a general account of how people decide where to move their eyes during the reading of Chinese (Liu et al., submitted for review). Thus, our variant Landolt-C paradigm has provided new information that is consistent with prior research using this paradigm (e.g., Williams & Pollatsek, 2007; Vanyukov et al., 2012) and with prior research on Chinese reading (e.g., Li et al., 2014; Liversedge et al., 2013; Yan et al., 2006), lending further support to a theoretical account of eye-movement control in Chinese reading—an account based on local character- and word-based control of when the eyes move that in turn affects where (or more precisely, how far) the eyes move (Liu et al., 2014, submitted for review; Wei et al., 2013). This represents significant empirical and theoretical progress in understanding eye-movement control during reading because it reveals the limitations of existing accounts of eye-movement control (e.g., the computational models cited earlier) that are based of years of research of alphabetic languages. It goes without saying that this progress will be extremely useful for developing formal accounts (i.e., computational models) of eye-movement control in non-alphabetic languages like Chinese, and more generally for informing our understanding of how 18 differences among languages and writing systems influence the perceptual, cognitive, and motoric processes that move the eyes during reading. 19 References Anderson, J. R. (1990). The adaptive character of thought. Hillsdale, NJ: Erlbaum. Barr, D. J., Levy, R., Scheepers, C., & Tily, H. J. (2013). Random effects structure for confirmatory hypothesis testing: Keep it maximal. Journal of Memory and Language, 68, 255-278. Bates, D., Maechler, M., Bolker, B., & Walker, S. (2014). lme4: Linear mixed-effects models using Eigen and S4. URL http://lme4.r-forge.r-project.org/. Bicknell, K., & Levy, R. (2012).The utility of modelling word identification from visual input within models of eye movements in reading. Visual Cognition, 20, 422-456. Chen, H. (1996). Chinese reading and comprehension: A cognitive psychology perspective. In M. H. Bond (Ed.), Handbook of Chinese psychology (pp. 43- 62). Hong Kong: Oxford University Press. Chen, H., Song, H., Lau, W. Y., Wong, K. F. E., & Tang, S. L. (2003). Chinese reading and comprehension: A cognitive psychology perspective. In C. McBride- Chang & H. Chen (Eds.), Reading development in Chinese children (pp. 157- 169). Westport, CT: Praeger. Chen, H., & Zhou, X. (1999). Processing East Asian languages: An introduction. Language and cognitive processes, 14, 425-428. Cheng, C. (1981). Perception of Chinese character. Act Psychological Taiwanica, 23, 137-153. Corbic, D., Glover, L., & Radach, R. (2007). The Landolt-C string scanning task as a proxy for visuomotor processing in reading. A pilot study. In Poster session presented at the 14th European Conference on Eye Movements. 20 Engbert, R., Nuthmann, A., Richter, E., & Kliegl, R. (2005). SWIFT: A dynamical model of saccade generation during reading. Psychological Review, 112, 777- 813. Feng, G. (2008). Orthography and eye movements: The paraorthographic linkage hypothesis. In K. Rayner, D. Shen, X. Bai, & G. Yan (Eds.), Cognitive and cultural influences on eye movements (pp. 395-420). Tianjin, China: Tianjin People’s Publishing House. Forster, K. I. & Chambers, S. M. (1973). Lexical access and naming time. Journal of Verbal Learning and Verbal Behavior, 12, 627-635. Henderson, J. M., & Ferreira, F. (1990). Effects of foveal processing difficulty on the perceptual span in reading: Implications for attention and eye movement control. Journal of Experimental Psychology: Learning, Memory, and Cognition, 16, 417-429. Hoosain, R. (1991). Aspects of the Chinese language. In R. Hoosain (Ed.), Psycholinguistic implications for linguistic relativity: A case study of Chinese (pp. 5–21). Hillsdale, NJ: Erlbaum. Hoosain, R. (1992). Psychological reality of the word in Chinese. In H. C. Chen & O. J. L. Tzeng (Eds.), Language processing in Chinese (pp. 111-130). Amsterdam, the Netherlands: North-Holland. Inhoff, A. W., & Rayner, K. (1986). Parafoveal word processing during eye fixations in reading: Effects of word frequency. Perception & Psychophysics, 40, 431- 439. Kennison, S. M., & Clifton, C., Jr. (1995). Determinants of parafoveal preview benefit in high and low working memory capacity readers: implications for eye 21 movement control. Journal of Experimental Psychology: Learning, Memory and Cognition, 21, 68-81. Kliegl, R., Nuthmann, A., & Engbert, R. (2006). Tracking the mind during reading: The influence of past, present, and future words on fixation durations. Journal of Experimental Psychology: General, 135, 12-35. Kuznetsova, A., Brockhoff, P. B., & Christensen, R. H. B. (2013). lmerTest: Tests for random and fixed effects for linear mixed effect models (lmer objects of lme4 package). URL http://lmertest.r-forge.r-project.org/. Li, X., Bicknell, K., Liu, P., Wei, W., & Rayner, K. (2014). Reading Is Fundamentally Similar Across Disparate Writing Systems: A Systematic Characterization of How Words and Characters Influence Eye Movements in Chinese Reading. Journal of Experimental Psychology: General, 143, 895-913. Li, X., Gu, J., Liu, P., & Rayner, K. (2013). The advantage of word-based processing in Chinese reading: Evidence from eye movements. Journal of Experimental Psychology: Learning, Memory, and Cognition, 39, 879-889. Li, X., Liu, P., & Rayner, K. (2011). Eye movement guidance in Chinese reading: Is there a preferred viewing location? Vision Research, 51, 1146-1156. Li, X., & Logan, G. (2008). Object-based attention in Chinese readers of Chinese words: Beyond Gestalt principles. Psychonomic Bulletin & Review, 15, 945- 949. Li, X., Rayner, K., & Cave, K. R. (2009). On the segmentation of Chinese words during reading. Cognitive Psychology, 58, 525-552. Liu, P., Li, W., Lin, N., & Li, X. (2013). Do Chinese readers follow the National Standard Rules for word segmentation during reading? PLoS One, 8, e55440. 22 Liu, Y., Li, X., & Pollatsek, A. (2014). Cognitive Control of Saccade Amplitude during the Reading of Chinese: A Theoretical Analysis and New Evidence. Manuscript submitted to review. Liu, Y., & Reichle, E. D. (2010). The emergence of adaptive eye movements in reading. In S. Ohlsson & R. Catrabone (Eds.), Proceedings of the 32nd Annual Conference of the Cognitive Science Society (pp. 1136-1141). Austin, TX: Cognitive Science Society. Liu, Y., Reichle, E. D., & Gao, D.-G. (2013). Using reinforcement learning to examine dynamic attention allocation during reading. Cognitive Science, 37, 1507-1540. Liu, Y., & Reichle, E. D., & Li, X. (2014). Parafoveal Processing Affects Outgoing Saccade Length During the Reading of Chinese. Journal of Experimental Psychology: Learning, Memory, and Cognition. Manuscript in press. Liversedge, S.P., Hyönä, J., & Rayner, K. (2013). Eye movements during Chinese reading. Journal of Research in Reading, 36, S1-S3. Ma, G., Li, X., & Pollatsek, A. (in press). There is no relationship between the preferred viewing location and word segmentation in Chinese reading. Visual Cognition. doi: 10.1080/13506285.2014.1002554. McConkie, G. W., Kerr, P. W., Reddix, M. D., & Zola, D. (1988). Eye movement control during reading: I. The location of the initial eye fixations on words. Vision Research, 28, 1107-1118. O’Regan, J. K. (1981). The convenient viewing position hypothesis. In D. F. Fisher, R. A. Monty, & J. W. Senders (Eds.), Eye movements: Cognition and visual perception (pp. 363–383). Hillsdale, NJ: Lawrence Erlbaum Associates. 23 Pinheiro, J.C., & Bates D.M. (2000). Mixed-effects Models in S and S-PLUS. Springer, New York, USA. Rayner, K. (1979). Eye guidance in reading: Fixation locations within words. Perception, 8, 21-30. Rayner, K. (1998). Eye movements in reading and information processing: 20 years of research. Psychological Bulletin, 124, 372-422. Rayner, K. (2009). Eye movements and attention in reading, scene perception, and visual search. Quarterly Journal of Experimental Psychology, 62, 1457-1506. Rayner, K., Ashby, J., Pollatsek, A., & Reichle, E. D. (2004). The effects of frequency and predictability on eye fixations in reading: Implications for the E-Z Reader model. Journal of Experimental Psychology: Human Perception and Performance, 30, 720-730. Rayner, K., Li, X., Juhasz, B. J., & Yan, G. (2005). The effect of word predictability on the eye movements of Chinese readers. Psychonomic Bulletin & Review, 12, 1089-1093. Rayner, K., & Morrison, R. E. (1981). Eye movements and identifying words in parafoveal vision. Bulletin of the Psychonomic Society, 17, 135-138. Rayner, K., Sereno, S. C., & Raney, G. E. (1996). Eye movement control in reading: A comparison of two types of models. Journal of Experimental Psychology: Human Perception and Performance, 22, 1188-1200. Reichle, E. D. (2011). Serial attention models of reading. In S. P. Liversedge, I. D. Gilchrist, & S. Everling (Eds.), Oxford handbook on eye movements (pp. 767- 786). Oxford, U.K.: Oxford University Press. 24 Reichle, E. D. & Laurent, P. (2006). Using reinforcement learning to understand the emergence of “intelligent” eye-movement behavior during reading. Psychological Review, 113, 390-408. Reichle, E. D., Pollatsek, A., Fisher, D. L., & Rayner, K. (1998). Toward a model of eye movement control in reading. Psychological Review, 105, 125-157. Reichle, E. D., Warren, T., & McConnell, K. (2009). Using E-Z Reader to model the effects of higher-level language processing on eye movements during reading. Psychonomic Bulletin & Review, 16, 1-21. Schad, D.J., & Engbert, R. (2012). The zoom lens of attention: Simulating shuffled versus normal text reading using the SWIFT model. Visual Cognition, 20, 391- 421. Schilling, H. E. H., Rayner, K., & Chumbley, J. I. (1998). Comparing naming, lexical decision, and eye fixation times: Word frequency effects and individual differences. Memory and Cognition, 26, 1270-1281. Smith, B. (1988). Gestalt Theory: An Essay in Philosophy. In B. Smith (Ed.), Foundations of Gestalt Theory (pp. 11-81), Munich and Vienna: Philosophia Verlag. Tsai, J. L., & McConkie, G. W. (2003). Where do Chinese readers send their eyes? In R. R. J. Hyona & H. Deubel (Eds.), The mind’s eye: Cognitive and applied aspects of eye movement research (pp. 159-176). Amsterdam, the Netherlands: Elsevier. Vanyukov, P. M., Warren, T., Wheeler, M. E., & Reichle, E. D. (2012). The emergence of frequency effects in eye movements. Cognition, 123, 185-189. Wei, W., Li, X., & Pollastsek, A. (2013). Word properties of a fixated region affect outgoing saccade length in Chinese reading. Vision Research, 80, 1-6. 25 Williams, C. C., & Pollatsek, A. (2007). Searching for an O in an array of Cs: Eye movements track moment-to-moment processing in visual search. Perception and Psychophysics, 69, 372-381. Williams, C. C., Pollatsek, A., & Reichle, E. D. (2014). Examing eye movements in visual search through clusters of objects in a circular array. Journal of Cognitive Psychology, 26, 1-14. Yan, G., Tian, H., Bai, X., & Rayner, K. (2006). The effect of word and character frequency on the eye movements of Chinese readers. British Journal of Psychology, 97, 259-268. Yan, M., Kliegl, R., Richter, E., Nuthmann, A., & Shu, H. (2010). Flexible saccade target selection in Chinese reading. The Quarterly Journal of Experimental Psychology, 63, 705-725. Yang, H., & McConkie, G. W. (1999). Reading Chinese: Some basic eye-movement characteristics. In J. Wang, A. W. Inhoff, & H.-C. Chen (Eds.), Reading Chinese script (pp. 207-222). Mahwah, NJ: Erlbaum. Zang, C., Liversedge, S.P., Bai, X., & Yan, G. (2011). Eye movements during Chinese reading. In S.P. Liversedge, I.D. Gilchrist, & S. Everling. (Eds). Oxford Handbook on Eye Movements (pp. 961-978). Oxford University Press. 26 Footnote 1. For the purposes of brevity and to draw parallels between our paradigm and the actual reading of Chinese text (which is of theoretical interest), we will refer to the Landolt-squares used in our experiment as “characters” and to the clusters of such characters as “words” throughout the remainder of this article. (Accordingly, we will also drop the use of scare quotes.) 27 Table 1. Chi-square tests for the uniformity of all forward fixation locations as a function of word size (i.e., number of characters). Word Size 1-character 2-character 3-character 4-character χ2 0.08 1.19 0.92 1.74 df 1 3 5 7 p 0.778 0.755 0.969 0.973 Note: The null hypothesis is uniformity of all forward fixation locations. 28 Table 2. LMM analyses of all forward fixation locations (aligned to the left of fixated word, in characters). Variables Intercept Practice (# of trials) Word N-1 # of characters gap size (pixels) Word N # of characters gap size (pixels) Word N+1 # of characters gap size (pixels) b 0.01 0.0001 -0.01 -0.001 0.99 0.01 0.01 -0.01 Model SE 0.07 0.0001 0.01 0.01 0.01 0.01 0.01 0.01 t 0.13 0.40 -1.20 -0.15 97.27 1.01 0.92 -1.24 p 0.894 0.689 0.232 0.878 < 0.001 0.315 0.358 0.214 - - 2.47 2.46 Min. 2.48 2.47 Values (characters) Max. 2.45 2.46 2.45 2.48 2.48 2.45 0.99 2.45 3.95 2.48 Notes: For number of characters, min. = 1 character, max. = 4 characters; for gap size, min. = 2 pixels, max. = 8 pixels; for practice, min. = 1st trial, max. = 256th trial. Estimates of predicted variable values were calculated while fixing the values of the other variables equal to their mean values. 29 Table 3. LMM analyses of forward saccade length (in characters). Variables Intercept Practice (# of trials) Word N-1 # of characters gap size (pixels) Word N # of characters gap size (pixels) Word N+1 # of characters gap size (pixels) b 2.88 0.0001 0.01 0.01 0.10 0.01 0.04 0.02 SE 0.19 0.0004 0.01 0.003 0.01 0.01 0.01 0.003 Model t 15.28 0.16 2.43 1.72 9.88 2.23 5.05 5.58 P < 0.001 0.877 0.015 0.085 < 0.001 0.037 < 0.001 < 0.001 - - 3.44 3.42 Min. 3.40 3.40 Values (characters) Max. 3.36 3.37 3.44 3.43 3.48 3.47 3.27 3.38 3.56 3.46 Notes: For number of characters, min. = 1 character, max. = 4 characters; for gap size, min. = 2 pixels, max. = 8 pixels; for practice, min. = 1st trial, max. = 256th trial. Estimates of predicted variable values were calculated while fixing the values of the other variables equal to their mean values. 30 Table 4. LMM analyses of first-fixation durations (ms). Model Variables Intercept Practice (# of trials) Word N-1 # of characters gap size (pixels) Word N # of characters gap size (pixels) Word N+1 # of characters gap size (pixels) b 319.70 -0.03 0.63 -0.07 -1.35 -1.16 1.46 -0.07 SE 10.80 0.02 0.49 0.34 0.67 0.33 0.93 0.32 t 29.61 -1.64 1.28 -0.20 -2.01 -3.50 1.57 -0.23 p < 0.001 0.117 0.200 0.847 0.051 0.002 0.131 0.819 - 308 - 315 310 312 Values (ms) Min. Max. 309 312 313 311 312 311 313 315 309 308 Notes: For number of characters, min. = 1 character, max. = 4 characters; for gap size, min. = 2 pixels, max. = 8 pixels; for practice, min. = 1st trial, max. = 256th trial. Estimates of predicted variable values were calculated while fixing the values of the other variables equal to their mean values. 31 Table 5. LMM analyses of gaze durations (ms). Variables Intercept Practice (# of trials) Word N-1 # of characters gap size (pixels) Word N # of characters gap size (pixels) Word N+1 # of characters gap size (pixels) Model t 8.68 0.59 p < 0.001 0.560 -3.34 -1.37 < 0.001 0.187 b 246.30 0.04 -4.03 -1.39 SE 28.39 0.07 1.21 1.02 135.70 -9.75 10.91 0.72 12.44 < 0.001 -13.47 < 0.001 -6.84 0.55 1.53 0.60 -4.46 0.92 < 0.001 0.357 - 528 - 517 529 527 Values (ms) Min. Max. 533 522 513 525 517 519 322 552 730 494 Notes: For number of characters, min. = 1 character, max. = 4 characters; for gap size, min. = 2 pixels, max. = 8 pixels; for practice, min. = 1st trial, max. = 256th trial. Estimates of predicted variable values were calculated while fixing the values of the other variables equal to their mean values. 32 Table 6. LMM analyses of total-viewing times (ms). Model Variables Intercept Practice (# of trials) Word N-1 # of characters gap size (pixels) Word N # of characters gap size (pixels) Word N+1 # of characters gap size (pixels) b 306.80 -0.08 -5.06 -1.66 158.70 -13.16 -6.53 -0.42 SE 31.93 0.08 1.41 1.11 3.66 0.85 1.41 0.70 t 9.61 -1.00 -3.59 -1.49 p < 0.001 0.328 < 0.001 0.151 43.37 < 0.001 -15.44 < 0.001 -4.62 -0.60 < 0.001 0.548 - 576 - 598 595 592 Values (ms) Min. Max. 597 589 578 586 580 582 349 627 825 548 Notes: For number of characters, min. = 1 character, max. = 4 characters; for gap size, min. = 2 pixels, max. = 8 pixels; for practice, min. = 1st trial, max. = 256th trial. Estimates of predicted variable values were calculated while fixing the values of the other variables equal to their mean values. 33 Figure Caption Figure 1. Examples of two Chinese sentences and their respective translations. In the first, the sequence of four underlined characters correspond to two words. In the second, the same characters correspond to a single word. Figure 2. Examples of experimental materials, with the target (i.e., a square) and one of the “word” clusters being rendered in gray for illustrative purposes. Figure 3. The distribution of initial fixations on words that received one or more fixations (panels a-b), of first-of-multiple fixations on words that received two or more fixations (panels c-d), of single fixations on words that received only one fixation (panels e-f), and of all forward fixations on words during first-pass scanning (panels g-h). All of the panels show the fixation-location distributions as a function of the number of characters within a word. Panels a, c, e, and g (left column) also show the fixation-location distributions as a function of the gap size within a word, whereas panels b, d, f, and g (right column) also show the fixation-location distributions as a function of practice (ordinal block number). Error bars represented the standard errors of the means. 34 Figure 1. 35 Figure 2. 36 Figure 3. 37
1511.00235
1
1511
2015-11-01T11:46:33
Broadband Macroscopic Cortical Oscillations Emerge from Intrinsic Neuronal Response Failures
[ "q-bio.NC", "cond-mat.stat-mech", "physics.bio-ph" ]
Broadband spontaneous macroscopic neural oscillations are rhythmic cortical firing which were extensively examined during the last century, however, their possible origination is still controversial. In this work we show how macroscopic oscillations emerge in solely excitatory random networks and without topological constraints. We experimentally and theoretically show that these oscillations stem from the counterintuitive underlying mechanism - the intrinsic stochastic neuronal response failures. These neuronal response failures, which are characterized by short-term memory, lead to cooperation among neurons, resulting in sub- or several- Hertz macroscopic oscillations which coexist with high frequency gamma oscillations. A quantitative interplay between the statistical network properties and the emerging oscillations is supported by simulations of large networks based on single-neuron in-vitro experiments and a Langevin equation describing the network dynamics. Results call for the examination of these oscillations in the presence of inhibition and external drives.
q-bio.NC
q-bio
Broadband Macroscopic Cortical Oscillations Emerge from Intrinsic Neuronal Response Failures Amir Goldental1,†, Roni Vardi2,†, Shira Sardi1,2, Pinhas Sabo1 and Ido Kanter1,2,* 1Department of Physics, Bar-Ilan University, Ramat-Gan 52900, Israel 2Gonda Interdisciplinary Brain Research Center and the Goodman Faculty of Life Sciences, Bar- Ilan University, Ramat-Gan, Israel †These authors contributed equally to this work. *Correspondence: [email protected] Abstract Broadband spontaneous macroscopic neural oscillations are rhythmic cortical firing which were extensively examined during the last century, however, their possible origination is still controversial. In this work we show how macroscopic oscillations emerge in solely excitatory random networks and without topological constraints. We experimentally and theoretically show that these oscillations stem from the counterintuitive underlying mechanism - the intrinsic stochastic neuronal response failures. These neuronal response failures, which are characterized by short-term memory, lead to cooperation among neurons, resulting in sub- or several- Hertz macroscopic oscillations which coexist with high frequency gamma oscillations. A quantitative interplay between the statistical network properties and the emerging oscillations is supported by simulations of large networks based on single-neuron in-vitro experiments and a Langevin equation describing the network dynamics. Results call for the examination of these oscillations in the presence of inhibition and external drives. 1. INTRODUCTION The most widespread cooperative activity of neurons within the cortex is spontaneous macroscopic oscillations(Silva et al., 1991;Gray, 1994;Contreras et al., 1997;Buzsaki and Draguhn, 2004;Chialvo, 2010), which range between sub- and hundred- Hertz(Başar et al., 2001;Brovelli et al., 2004;Buzsaki and Draguhn, 2004;Grillner et al., 2005;Giraud and Poeppel, 2012). The high cognitive functionalities of these oscillations are still controversial(Klimesch, 1996;1999;Başar et al., 2001;Wiest and Nicolelis, 2003;Buzsaki, 2006;Kahana, 2006;Fries, 2009;Giraud and Poeppel, 2012;Thivierge et al., 2014) and are typically attributed to transitory binding activities among indirect macroscopic distant cortical regions(Gray, 1994;Başar et al., 2001;Buzsaki and Draguhn, 2004;Roxin et al., 2004;Fries, 2009). In Cooperative oscillations emerge from response failures addition, it was found that the theta rhythms(Klimesch, 1999;Buzsaki and Draguhn, 2004), oscillations in the range of 4-10 Hz, play a key role in the formation and retrieval of episodic and spatial memory(Hasselmo, 2005). This theta rhythm is usually accompanied by high frequency oscillations in the range of 30-80 Hz, known as gamma oscillations(Colgin and Moser, 2010). Gamma oscillations are also related to sensory stimulations and induce neuronal ensemble synchrony by generating a narrow window for effective excitation(Cardin et al., 2009). There are several suggested mechanisms for the formation of such rhythms on the network level(Wang, 2010). Most of the proposed mechanisms are based on the existence of inhibitory synapses(Wilson and Cowan, 1972;Jirsa and Haken, 1996;Brunel and Wang, 2003), especially for high frequency oscillations(Brunel and Wang, 2003;Colgin and Moser, 2010;Wang, 2010). For illustration, assume that a fast excitation increases neural firing in an excitatory short-delayed feedback loop. Consequently, neuronal populations along the excitatory feedback loop will fire at high rates that will cause a slower response of the inhibitory neurons. As a result, the inhibitory neurons will depress the activity within the excitatory population. This will then depress the excitation of the inhibitory population. Finally, the depression of the inhibitory neurons allows a repeated fast excitation of the excitatory population. In this work we show how extra-cellular potential oscillations, synchronized rhythmic firing of neurons, emerge in random networks without inhibitory synapses. Our findings are based on an experimental observations of neuronal plasticity in the form of intrinsic neuronal response failures(Vardi et al., 2015). Using simulations of large networks, based on single-neuron in-vitro experiments, we show that this type of neuronal plasticity leads to the coexistence of both theta and gamma oscillations. Results are supported by a quantitative approach based on a Langevin equation, which describes the network dynamics. 2. MATERIALS AND METHODS 2.1 In-vitro experiments 2.1.1 Animals All procedures were conducted in accordance with the National Institutes of Health Guide for the Care and Use of Laboratory Animals and Bar-Ilan University Guidelines for the Use and Care of Laboratory Animals in Research and were approved and supervised by the Institutional Animal Care and Use Committee. 2.1.2 Culture preparation Cortical neurons were obtained from newborn rats (Sprague-Dawley) within 48 h after birth using mechanical and enzymatic procedures. The cortical tissue was digested enzymatically with 0.05% trypsin solution in phosphate-buffered saline (Dulbecco’s PBS) free of calcium and magnesium, and supplemented with 20 mM glucose, at 37◦C. Enzyme treatment was terminated using heat-inactivated horse serum, and cells were then mechanically dissociated. The neurons were plated directly onto substrate-integrated multi- 2 Cooperative oscillations emerge from response failures electrode arrays (MEAs) and allowed to develop functionally and structurally mature networks over a time period of 2-3 weeks in-vitro, prior to the experiments. Variability in the number of cultured days in this range had no effect on the observed results. The number of plated neurons in a typical network was in the order of 1,300,000, covering an area of about 380 mm2. The preparations were bathed in minimal essential medium (MEM-Earle, Earle's Salt Base without L-Glutamine) supplemented with heat-inactivated horse serum (5%), glutamine (0.5 mM), glucose (20 mM), and gentamicin (10 g/ml), and maintained in an atmosphere of 37◦C, 5% CO2 and 95% air in an incubator as well as during the electrophysiological measurements. 2.1.3 Synaptic blockers All experiments were conducted on cultured cortical neurons that were functionally isolated from their network by a pharmacological block of glutamatergic and GABAergic synapses. For each culture 20 l of a cocktail of synaptic blockers was used, consisting of 10 μM CNQX (6-cyano-7-nitroquinoxaline-2,3- dione), 80 μM APV (amino-5-phosphonovaleric acid) and 5 μΜ bicuculline. This cocktail did not block the spontaneous network activity completely, but rather made it sparse. At least one hour was allowed for stabilization of the effect. 2.1.4 Stimulation and recording An array of 60 Ti/Au/TiN extracellular electrodes, 30 μm in diameter, and spaced 500 μm from each other (Multi-Channel Systems, Reutlingen, Germany) was used. The insulation layer (silicon nitride) was pre- treated with polyethyleneimine (0.01% in 0.1 M Borate buffer solution). A commercial setup (MEA2100- 2x60-headstage, MEA2100-interface board, MCS, Reutlingen, Germany) for recording and analyzing data from two 60-electrode MEAs was used, with integrated data acquisition from 120 MEA electrodes and 8 additional analog channels, integrated filter amplifier and 3-channel current or voltage stimulus generator (for each 60 electrode array). Mono-phasic square voltage pulses typically in the range of [-800, -500] mV and [60, 400] μs were applied through extracellular electrodes. Each channel was sampled at a frequency of 50k samples/s, thus the changes in the neuronal response latency were measured at a resolution of 20 s. 2.1.5 Cell selection Each node was represented by a stimulation source (source electrode) and a target for the stimulation – the recording electrode (target electrode). These electrodes (source and target) were selected as the ones that evoked well-isolated, well-formed spikes and reliable response with a high signal-to-noise ratio. This examination was done with a stimulus intensity of -800 mV with duration of 200 μs using 30 repetitions at a rate of 5 Hz, followed by 1200 repetitions at a rate of 10 Hz. 2.1.6 In-vitro experiment with feedback loops and neural circuits The activity of all source and target electrodes was collected and action potentials were detected on-line by threshold crossing, and entailed stimuli were delivered in accordance with the circuit's connectivity, as 3 Cooperative oscillations emerge from response failures described below. A successful response was defined as a spike occurring within a typical time window of 2-10 ms following the beginning of an electrical stimulation. In Figures 2A,B, after every spike detection two supra-threshold extracellular stimulations were given to the same neuron, after 600 ms and 630 ms. In case that the timings of the stimulations overlap, only one stimulation is given. In Figures 2C,D, after every spike detection supra-threshold extracellular stimulations were given to its connected neurons. For example, if a spike was detected at the left (green) neuron (Figures 2C), a supra- threshold extracellular stimulation will be given to the middle (brown) neuron after 330 ms. 2.1.7 Data analysis Analyses were performed in a Matlab environment (MathWorks, Natwick, MA, USA). The reported results were confirmed based on at least eight experiments each, using different sets of neurons and several tissue cultures. The temporal firing frequency, around stimulation no. i, of the neuron in Figure 1 was calculated using the following procedure 𝑓𝑖𝑟𝑖𝑛𝑔 𝑓𝑟𝑒𝑞𝑢𝑒𝑛𝑐𝑦 (𝑖) = 𝑠𝑡𝑖𝑚𝑢𝑙𝑎𝑡𝑖𝑜𝑛 𝑓𝑟𝑒𝑞𝑢𝑒𝑛𝑐𝑦 ∙ ∑ 𝑖+125 m=max (0,𝑖−125) 𝐼𝑠_𝑆𝑝𝑖𝑘𝑒𝑑(𝑚) 𝑖 + 125 − max(0, 𝑖 − 125) where Is_Spiked(m) = 1 if the neuron responded to stimulation no. m, otherwise Is_Spiked(m) = 0. 2.2 Simulations Simulations (similar to (Vardi et al., 2015)) consist of a network of N leaky integrate and fire neurons 𝑑𝑉𝑖 𝑑𝑡 = − 𝑉𝑖 𝜏 + ∑ 𝑁 𝑗=1 𝐽𝑗𝑖 ∑ 𝛿(𝑡 − 𝑡′ − 𝐷𝑗𝑖) 𝑡′ where i∈[1,N], =20 ms, Jji and Dji are the connection’s strength and delay from neuron j to i, respectively. The summation over t' sums all firing times of neuron j, the integration time step is 0.05 ms, and the threshold is 1. For the nth threshold crossing of a neuron, its probability for a response is [Σ(n-m/c)exp(- m)]/[Σexp(-m)], where n is the time gap between the nth and (n-1)th threshold crossings, c=1/fc, =1.4 and the sum is over m≥0. A refractory period of 2 ms is imposed after an evoked spike, for response failures the voltage is set to 0.2. Results were found to be insensitive to initial conditions. 2.2.1 Various forms of p(s) lead to the same <ISI> The probability for a response, given the last inter-stimulation-interval, p(s), should lead to <ISI>=c (Figures 1 and 3). One can show that any p(s) satisfying ∞ ∫ 𝑝(𝑠𝜏)𝑝(𝜏)𝑑𝜏 0 ∞ ∫ 𝜏𝑝(𝜏)𝑑𝜏 0 = 1 𝜏𝑐 4 Cooperative oscillations emerge from response failures where p() is the probability density of an inter-stimulation-interval equals to  leads to <ISI>=c. The numerator on the left hand side stands for the average probability for a successful response, and the denominator stands for the average inter-stimulation-interval. This ratio is equivalent to the firing rate, hence equals to 1/c. For instance p(s)=/c, this theoretical curve fits all p() (Figure 3D). In the activity of some random networks a good approximation for p() is 𝑝(𝜏) = 2 𝜏𝑐 𝑒𝑥𝑝 (− 2𝜏 𝜏𝑐 ). For this p() some of the p(s) solutions, which lead to <ISI>=c, are p(s)=0.5, p(s)=/c and 𝑝(𝑠𝜏) = 1 − 𝑒𝑥𝑝 (− 2𝜏 𝜏𝑐 ), which is similar to Figure 3D. 2.2.2 Fourier analysis of the rate To perform a Fourier analysis on the activity of the network we define the rate vector: 𝑅(𝑡𝑖𝑚𝑒 = 𝑖 ∙ 𝑤) = 1 𝑁𝑤 (𝑖+0.5)𝑤 ∫ (𝑖−0.5)𝑤 ∑ 𝛿(𝑡′ − 𝑡𝑠𝑝𝑖𝑘𝑒) 𝑡𝑠𝑝𝑖𝑘𝑒 𝑑𝑡′ where i is an integer, w is a predefined time window and the sum is over all spike times of all N neurons. Next, a discrete Fourier transform is preformed and the function of resulting amplitudes is normalized and smoothed using a sliding window of 1 Hz. 3. RESULTS 3.1 Neuronal plasticity: Neuronal response failures We start with the quantification of the neuronal response latency (NRL)(Vardi et al., 2013a;Marmari et al., 2014;Vardi et al., 2015), measured as the time-lag between a stimulation and its corresponding evoked spike (Figure 1). It was recently shown(Vardi et al., 2015) that when a neuron is repeatedly stimulated, its NRL stretches gradually (Figure 1, upper panel), and when the stimulation frequency is high enough stochastic neuronal response failure (NRFs) emerge. Specifically, each neuron is characterized by a critical frequency, fc, typically ranging among neurons between 2 and 30 Hz. Stimulation frequencies above fc result in NRFs, whereas for supra-threshold stimulations below fc a response is assured. The probability of the NRFs is such that the neuron functions similar to a low pass filter, saturating its firing rate (Figure 1, lower panel). Quantitatively, for a stimulation frequency f, the NRF probability is 1-fc/f, i.e. the firing frequency is saturated at fc. Thus, changing the stimulation frequency will change the probability for NRFs, while the firing frequency remains bounded, fc. This observation is demonstrated using a cultured cortical neuron, functionally separated from its network by synaptic blockers, with above-threshold stimulations (see section 2.1 In-vitro experiments). We examine a neuron with periodic stimulation trials of 10, 12 and 15 Hz, and NRFs appear after a short transient where 5 Cooperative oscillations emerge from response failures the neuron exhibits an increase of its NRL (Figure 1, upper panel). Examining the temporal firing rate of the neuron (Figure 1, lower panel), it is noticeable that the firing frequency of the neuron is saturated at fc=5 Hz, independent of the stimulation frequency. The effect of NRFs is first examined on small neuronal circuits using the following experiment: We stimulate the neuron ones and impose on the neuron two self-feedback delay loops, e.g. 600 ms and 630 ms (Figure 2A). The neuron is stimulated 600 ms and 630 ms after each evoked spike (see section 2.1.6 In- vitro experiment with feedback loops and neural circuits). In the case of vanishing probability for NRF, the neuron should fire every 30 ms (Kanter et al., 2011;Vardi et al., 2012b), i.e. 33.3 Hz. Since fc=5 Hz, some NRFs appear (Figure 2B) forming bunched firing, separated by ~600 ms, whereas the intra-bunch time- lags, 30 ms, originated from the difference between the two feedback delay loops (Figure 2B). The emergence of such firing bunches indicate some dynamical changes of the NRF’s probability, where the neuron adapts its failure probability as a result of its recent stimulation history. A more biologically realistic scenario is a neuronal circuit consisting of three artificially connected neurons (Vardi et al., 2012b), forming the same delay loops (Figure 2C), 600 ms and 630 ms. In addition, each neuron is identified by different fc and only the central neuron receives the initial stimulation. For the central neuron, which is characterized by fc=7.2 Hz, NRFs occur, since the frequency of its driven neurons is greater than its critical frequency, 6.8+4>7.2 Hz, and similarly the outer neurons have response failures (7.2>6.8, 7.2>4) (Figure 2D). Besides the formation of firing bunches for each neuron, their firing are correlated at zero- or shifted- time-lags (Figure 2D). The question whether the repeated bunches in such small neuronal circuits shed light on macroscopic cortical oscillations requires to sail towards large scale simulations. 3.2 Short-term memory of neuronal plasticity The observed firing bunches indicate a form of short-term memory of neuronal plasticity, where the NRF probability is mainly a function of the few preceding stimulations. Our next goal is to experimentally quantify this neuronal plasticity and then examine its implementation on the dynamics of large neural networks using large scale simulations. We first assume a simplified network where each neuron has two above-threshold inputs, two outputs and the same fc (Figure 3A), hence the statistics of the inter-stimulation-intervals of each neuron is expected to approximately follow an exponential distribution with 2fc rate, Exp(2fc) (Figure 3B, upper panel). To quantify the statistics of the NRFs, a long stimulation trail obeying Exp(2fc), underc/8 time resolution, was given to a cultured neuron with fc~5 Hz (Figure 3B). Next, the conditional probabilities for a successful response (an evoked spike), p(sii-1), were estimated for events where the current inter-stimulation- interval equals i and the previous one equals i-1 (Figures 3B,C). It is evident that p(sii-1) is primarily a function of i, i.e. the probability for a successful response is dictated mainly by the current inter- 6 Cooperative oscillations emerge from response failures stimulation-interval, i. Hence, the NRFs can be dynamically approximated using p(s) (Figure 3D), indicating that the neuronal response failure probability might be fairly estimated based solely on the last inter-stimulation-interval, . 3.3 Neuronal oscillations on the network level The experimental estimation of p(s) is utilized to simulate a large scale network (Figure 3A) and is exemplified for 2000 neurons where delays between connected neurons, D, are randomly selected from a uniform distribution U(10,15) ms. The simulation is initialized with timings of evoked spikes for a subset of the neurons, however, besides the transient time results were found to be insensitive to the initial conditions. The response failure is then randomly selected following the experimentally measured p(s), independently for each neuron and stimulation (Figure 3D). Indeed, the assumption Exp(2fc) (Figure 3B, upper panel) was confirmed (Figure 3E), the statistics of the inter-stimulation-intervals of each neuron approximately follows an exponential distribution with 2fc rate. The raster plot of the network firing as well as the time-dependent firing rate (Figures 4A,B) clearly indicate cooperative oscillation which can be quantified using the Fourier analysis to fosc~3 Hz (Figure 4C, see section 2.2), and are absent in the Fourier analysis of the firing of each individual neuron (Figure 3F). Another broadened Fourier peak is centered at f~80 Hz, gamma oscillations(Brunel and Wang, 2003;Cunningham et al., 2004;Fries, 2009;Minlebaev et al., 2011;Dugladze et al., 2012), which is attributed to the average delay, <D>=12.5 ms. It reflects the average firing frequency of each neuron where NRFs vanish and all delays are equal to <D>, as GCD=<D> for delay loops of such a random network(Kanter et al., 2011;Vardi et al., 2012a). Similar cooperative oscillations were obtained using a counterpart simulation for the same network (Figures 4D,F) while using the theoretical form p(s) (Figure 3D). Although the form of p(s) varies among neurons as well as between the theoretical and the experimental forms (Figure 3D), the cooperative oscillations are found quantitatively to be only slightly affected by its exact form. The robustness of fosc was also confirmed in simulations for more realistic scenarios where fc significantly varies among neurons as well as their input and output connectivity distributions. A more biological realization is exemplified in Figures 4G-I. The number of connections per neuron is much greater than 2, i.e. more than 50 pre- and 50 post- synaptic connections, where most of them are sub-threshold and on the average 1.5 of pre- and post-synaptic are above-threshold. Specifically, each sub-threshold connection produces an excitatory postsynaptic potential, EPSP, which is equal to 0.03 of the threshold. It is apparent that these additional connections do not qualitatively change the oscillations. In addition, f was found to be robust to a wider distribution of delays and followed its center (Figure 4I). Note that without these intrinsic NRFs, i.e. p(s)=1, the firing of each neuron is only bounded by the refractory period which is in the order of several milliseconds. In this limiting 7 Cooperative oscillations emerge from response failures case, the abovementioned theta and gamma oscillations disappear, as was confirmed in simulations (not shown). 3.4 Cortical oscillations versus statistical properties of the network The origin of the fast oscillations, f, (Figures 4C,F,I) is evident, however, the mechanism underling the slow cooperative oscillations(Wu et al., 1999;Sanchez-Vives and McCormick, 2000;Bollimunta et al., 2008;Nir et al., 2008;Crunelli and Hughes, 2010;Bollimunta et al., 2011), fosc, is still unclear. To explore this mechanism we identify the following two characteristic distances on the network. The first distance, Path, is the average minimal path between pairs of nodes, and the second distance, Loop, is the average over the minimal feedback loop of each node, both counted by the number of nodes along the route (Figure 5A). Numerical estimations based on various network topologies indicate that the average values of these two distances as well as their distributions are almost identical (Figure 5B) and their scaling decrease as 1/ln(<K>), where <K> is the average neuronal input connectivity (Figure 5C). These identities and scaling (Figures 5B,C), are also supported by the following theoretical argument. Assume a random network consisting of N neurons and an average connectivity <K>. The quantity Q(m) denotes the number of new connected neurons to a seed neuron at a distance of m neurons, hence Q is proportional to the probability (green) in Figure 5B. Start at an arbitrary neuron, Q(0)=1, this neuron is connected (pre-synaptic) to Q(1) neurons. Using a recursive formula one can approximate Q(i) = N (1 − exp (−< K > Q(i−1) N )) (1 − ∑ i−1 m=1 Q(m) N ) where N(1-exp(-<K>Q(i-1)/N)) stands for the average number of neurons at a distance i, which are connected from new neurons at distance i-1. The rightmost term is the probability that the neuron at distance i is a new neuron which was not counted at shorter distances, m<i. This recursive relation is solved numerically and the normalized Q(i) are presented in Figures 5B. The above analysis is valid for Loops and Paths, hence their statistics are identical, in agreement with the sampling of these quantities in Figures 5B. The importance of this distance in the formation of fosc can be understood according to the following argument. Assume a random subgroup of firing neurons activates another random group of neurons and vice versa. The minimal delay between pairs of neurons belonging to the two subgroups is Path·<D>. Consequently, the oscillations are expected to scale with Path·<D>, and indeed results indicate that fosc∝ln(<K>)∝(Path)-1 (Figures 5C,D). The minimal path is the most reliable one with respect to the NRFs, however, the effect of longer paths is not negligible as they might maintain the activity of the Path (Figure 2), especially as their entropy is higher. In addition, the NRFs are responsible to limit the firing of all the 8 Cooperative oscillations emerge from response failures network simultaneously (Figures 4B,E,H). Assume one neuron fires and activates <K> neurons after <D> ms, hence after m<D> ms, <K>m neurons fire. This exponential firing growth is bound by mPath, as self- feedback loops (Figures 2A,B) significantly lead to NRFs and to a fast decrease in the firing rate. In addition, for a given network topology, fosc is found to scale with ln(fc) (Figure 5E), and to be robust for networks composed of neurons with different fc (Figures 4G,H,I). These predictions might be realized in further experiments by controlling the network topology either by the neuronal concentration or by pharmacological manipulations. 3.5 Analytical description of the network oscillations An analytical description of fosc is also possible, and to simplify the presentation the method is briefly described for homogeneous networks only. Each node has the same fixed input and output connectivity, K, and all delays are equal to D ms, hence neurons can fire only at i·D ms, where i is an integer. The fraction, R(m), of neurons that fire at step m is given by 𝑅(𝑚) = 𝜒(𝑚) ∙ 𝐾 ∙ 𝑅(𝑚 − 1) + 𝜉(𝑚) (1) where (m) stands for the time-dependent white noise representing the stochastic nature of the response failures. The function (m) represents the susceptibility of the network, i.e. the fraction of neurons that fires if all neurons are stimulated at step m, and is explicitly given by ∞ 𝑖 𝜒(𝑚) = ∑ 𝑝(𝑠𝑖) ∙ 𝐾 ∙ 𝑅(𝑚 − (𝑖 + 1)) ∏(1 − 𝐾 ∙ 𝑅(𝑚 − 𝑛)) . (2) 𝑖=0 𝑛=2 The first term, p(si), stands for the probability for an evoked spike when the previous stimulation was given before i steps (Figure 3D). The term KR(m-(i+1)) pinpoints a neuron stimulated before i steps, and the product, the rightmost term, indicates that the neuron was not stimulated since step (m-i). Equation (2) indicates that (m) is a function of the variable R(l) only, with l<m. Hence, after the insertion of equation (2), into equation (1), one finds a recursion relation for R(m) which can be solved numerically given the initial conditions. The dynamical solution of this recursion relation revealed oscillations which were found to fit fairly good with those observed in large-scale simulations (Figures 4C,F,I). The equations imply that the network has some memory of its previous activity, which dictates the responsiveness of the entire network. This analytical description can be generalized to advanced structured networks, including random connectivity, distribution for the delays as well as to include variations among neuronal critical frequencies, fc. 4. Discussion We have demonstrated that intrinsic neuronal response failures drive a neural network activity towards oscillations, where high frequency oscillations, gamma, and low frequency oscillations, delta and theta, 9 Cooperative oscillations emerge from response failures coexist. The high frequency oscillations correspond to the average delay between connected neurons in the network, while low frequency oscillations are governed by statistical properties of the network, e.g. the average number of connections per neuron and the average critical frequency of neurons. The coexistence of high and low frequency oscillations was confirmed in a new type of simulations, based on a single neuron in-vitro experiments, to evaluate the firing activity of complex networks. Results were also supported by an analytical description of the stochastic dynamics of the network. Preliminary results in-vivo support our findings. The NRL increases by several milliseconds under periodic stimulations and terminates at an intermittent phase(Vardi et al., 2015). This phase is characterized by fluctuations around a constant NRL and accompanied by NRFs. Results indicate that fc can be below 10 Hz and vary among neurons. However, quantitative measurements of fc and the statistics of the NRFs require long stimulation trials, i.e. many thousands of high frequency periodic stimulations, which are currently beyond our experimental capabilities. The average firing rate of neurons in the network is low, e.g. ~3.6, ~3.9 and ~2.6 Hz in Figures 4B,E,H, respectively. These network low firing rates are lower than the neuronal critical frequency in Figures 4B,E, fc=5.7 Hz, and <fc>=6.5 Hz, in Figure 4H. Surprisingly, the neuronal critical firing frequency is not saturated even when the network is completely excitatory. A biological mechanism that suppresses the firing frequency of a single neuron below fc is aperiodic time-lags between stimulations (Vardi et al., 2015). For illustration, assume a slow mode of alternation between stimulation frequencies of 2fc (0.5c) and 2fc/3 (1.5c), such that <>=c. For the high and low frequency mode, the expected probability for a NRF is 0.5 and 0, respectively. Consequently, <ISI>=0.5(1.5c+c)=1.25c, corresponding to a lower firing rate, 0.8fC. In addition, the firing rates are considerably lower than f indicating that high frequency network oscillations consist of temporarily synchronized sub-groups of neurons. Indeed, the Fourier spectrum of a single neuron does not exhibit any dominant peaks (Figure 3F). Although the formation of broadband network oscillations is usually attributed to the existence of inhibition, it is evident that another possible mechanism is intrinsic neuronal response failures that dynamically drives neural networks to generate coherent oscillations with low averaged firing rates(Vardi et al., 2015). These observations raise the question of which functionalities demand synaptic inhibition. It was shown that inhibition slightly suppresses the network firing frequency even further(Vardi et al., 2015) and it also might change the amplitude of the oscillations. An additional possible hypothesis is that the role of inhibitory connections is to allow some neuronal computations which are based on conditional temporal formation of neuronal firing patterns. This type of functionality is an exclusive property of inhibitory synapses which probabilistically block an evoked spike of its driven nodes in a given time window(Vardi et al., 2013b;Goldental et al., 2014). In addition, the coexistence of the network oscillations with neuronal inhibition is intriguing, and especially the questions whether inhibition induces more modes of oscillations, 10 Cooperative oscillations emerge from response failures sharpens the existing ones, or suppresses the oscillatory behavior and stabilizes the network activity. Preliminary results of simulations indicate that inhibition might suppress the amplitude of the oscillations in the low frequency range and sharpen the oscillations in the gamma range. However, results might be sensitive to the selected parameters. The interplay between the presented spontaneous cortical oscillations and external stimulations given to the network is another intriguing question. Specifically, it is interesting to examine the coexistence and the interplay between the spontaneous oscillation frequencies determined by the network topology and the frequencies of the induced external stimulations. The understanding of these dynamics will shed light on the emerging cortical oscillations among coupled networks characterized by different statistical properties. ACKNOWLEDGEMENTS This research was supported by the Ministry of Science and Technology, Israel. The authors declare no competing financial interests. AUTHOR CONTRIBUTION R.V. and S.S. prepared and performed the experiments; R.V., S.S. and A.G. analysed the data; P.S. designed the experimental real-time interface; A.G. performed the simulations and developed the theoretical framework with help of P.S.; I.K., R.V. and A.G. wrote the manuscript. 11 Cooperative oscillations emerge from response failures FIGURE 1 Neuronal plasticity – in-vitro experiments. Upper panel: The NRL of a neuron stimulated at 10, 12 and 15 Hz (light blue, green and purple dots, respectively). Response failures are denoted by NRL<2 ms. Lower panel: The firing frequency calculated from the averaged ISI using sliding windows of 250 stimulations, or the maximal available one for Stim<250. 12 Cooperative oscillations emerge from response failures FIGURE 2 Firing bunches stem from neuronal response failures – in-vitro experiments. (A) Schematic of the neuron in Figure 1 characterized by fc=5 Hz, with 600 and 630 ms self-feedback loops. (B) A 3.5 s snapshot of the experimental results of (A) where stimulations (blue lines) and their corresponding evoked spikes (blue dashed lines) were recorded after an offset of to=21 s and a preparation at 5 Hz stimulation frequency over 300 s to settle the neuron at the intermittent phase. (C) Schematic of a circuit consisting of three different neurons with fc=6.8 (green), 7.2 (brown) and 4.0 (purple) Hz and 600 and 630 ms delay loops, similar to (A), (different neurons than the one in (A)). (D) A 5 s snapshot of the experimental results of (C) where ten initial stimulations were given to the central neuron (brown) at 4 Hz and to=25 s. Stimulations given to the colored neurons (brown/green/purple lines, respectively) and their corresponding evoked spikes (brown/green/purple dashed lines, respectively). A zoom-in of the gray area is presented (bottom). 13 Cooperative oscillations emerge from response failures FIGURE 3 The short-term memory of the stochastic neuronal response failures – in-vitro experiments and simulations. (A) Schematic of a prototypical examined excitatory network, where each neuron has two pre- and two post- synaptic connections and the same fc. Inset: The time-lags between neuronal stimulations is expected to approximately follow an exponential distribution, Exp(2fc). (B) Upper panel: The stimulation scheme where a neuron is stimulated under the resolution of c/8, such that the discrete differences between two consecutive stimulations, , follows Exp(2fc). Middle panel: Experimental NRL of a cultured neuron with fc~5 Hz under a long trial of stimulations following Exp(2fc), response failures are denoted at NRL=3.5 ms. Lower panel: Zoom-in of the middle panel (gray area) and schematic of the conditional probability p(sii-1), measuring the probability of a successful response, spike, given that the current inter-stimulation-interval equals i and the previous one equals i-1. (C) The probabilities 14 Cooperative oscillations emerge from response failures p(sii-1) obtained from the experimental data in (B) for time>3500 (middle panel). Each colored line presents p(sii-1) for a given i-1 in c/8 time units (legend). (D) The experimentally measured p(si) (black), measuring the probability of a successful response, spike, given that the current inter-stimulation- interval equals i, and the theoretically predicted one (green) using the simplified assumption, p(si)=i/c for i<c. For both curves, the average ISI~c is preserved. (E) The probability density function of inter- stimulation-intervals, , for all neurons of Figure 4A (blue). (F) Typical Fourier amplitude of spike timings of a randomly chosen neuron, taken from Figure 4A. 15 Cooperative oscillations emerge from response failures FIGURE 4 Cooperative cortical oscillations on a network level. (A) Raster plot of the evoked spikes (blue dots) obtained in the simulation of a network of 2000 neurons with fc=5.7 Hz. Each neuron has two randomly selected pre- and post- synaptic connections, and the simulation is based on the experimentally obtained p(s), (Figure 3D). Delays are randomly selected from U(10,15) ms. The contrast of the raster was enhanced using a dilution of a constant amount of randomly chosen points in each sliding window of 23 ms, with a step of 0.23 ms. The average dilution is ~60% of the points. (B) The average firing rate per neuron as a function of time, calculated for windows of 1 ms. (C) The normalized Fourier amplitude, using a sliding window of 1 Hz, of the entire firing of all neurons over a time slot of 30 s, indicating fosc~3.6 Hz and f~80 Hz. Inset: The normalized Fourier amplitudes in the range [0,30] Hz obtained from R(m), equation (1), D=12.5 ms (red) and from the simulation, ((B), blue). (D-F) Similar to (A-C) where each neuron has on average two pre- and post- randomly chosen synaptic connections and utilizing the theoretical p(s), (Figure 3D). fosc~4.0 Hz and f~80 Hz at (F). (G-I) Simulation of a network of 2000 neurons where each neuron has on the average 1.5 pre- and 1.5 post- synaptic above-threshold connections, and 50 pre- and 50 post- synaptic sub-threshold connections with a strength of 0.03, relative to a threshold of 1. p(si) 16 Cooperative oscillations emerge from response failures is generalized to an exponential decay function of the neuronal stimulation history, Σi-m/c)exp(- m)/[Σexp(-m)], =1.4 and the sum is over stimulation history, m≥0. Delays are randomly selected from U(12.5,20) ms and fc from U(3,10) Hz. fosc~2.6 Hz and f~60 Hz at (I), and the inset is similar to (C) and (F), but with K=1.5, fc=6.5 Hz and D=16.25 ms. 17 Cooperative oscillations emerge from response failures FIGURE 5 Scaling properties of fosc and f. (A) Schematic of an excitatory network where a self- feedback loop (light-red line) and the minimal self-feedback loop (red line) for a given neuron (filled red circle) are denoted. Similarly, a path (light-blue line) between two neurons (filled blue circles) and the minimal path (blue line) are denoted. (B) The distribution and its average (vertical lines) for the minimal path (Path) and for the minimal loop (Loop) obtained in simulations for networks as in (Figure 3A) with 18 Cooperative oscillations emerge from response failures N=4000, error bars are comparable with the circles. The analytical estimation is shown in green. (C) The scaling of the averaged quantities in (B) as a function of the average connectivity, <K>. (D) Simulation results indicate foscln<K>, where N=4000, fc is randomly chosen for each neuron from U(5,15) Hz and delays are randomly chosen from U(10,15) ms. The probability for a connection between two neurons is <K>/N. Error bars indicate the standard deviation. (E) Simulation results indicate foscln(fc), for networks as in (D), with <K>=2, but fc is the same for all neurons. (F) Simulation results indicate f1/<D>, for networks as in (D), with <K>=2, but delays are randomly chosen from U(<D>-2.5,<D>+2.5) ms. 19 Cooperative oscillations emerge from response failures REFERENCES Başar, E., Başar-Eroglu, C., Karakaş, S., and Schürmann, M. (2001). Gamma, alpha, delta, and theta oscillations govern cognitive processes. International Journal of Psychophysiology 39, 241-248. Bollimunta, A., Chen, Y., Schroeder, C.E., and Ding, M. (2008). Neuronal mechanisms of cortical alpha oscillations in awake-behaving macaques. The Journal of neuroscience 28, 9976-9988. Bollimunta, A., Mo, J., Schroeder, C.E., and Ding, M. (2011). Neuronal mechanisms and attentional modulation of corticothalamic alpha oscillations. The Journal of Neuroscience 31, 4935-4943. Brovelli, A., Ding, M., Ledberg, A., Chen, Y., Nakamura, R., and Bressler, S.L. (2004). Beta oscillations in a large-scale sensorimotor cortical network: directional influences revealed by Granger causality. Proceedings of the National Academy of Sciences of the United States of America 101, 9849-9854. Brunel, N., and Wang, X.-J. (2003). What determines the frequency of fast network oscillations with irregular neural discharges? I. Synaptic dynamics and excitation-inhibition balance. Journal of neurophysiology 90, 415-430. Buzsaki, G. (2006). Rhythms of the Brain. Oxford University Press. Buzsaki, G., and Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science 304, 1926-1929. Cardin, J.A., Carlén, M., Meletis, K., Knoblich, U., Zhang, F., Deisseroth, K., Tsai, L.-H., and Moore, C.I. (2009). Driving fast-spiking cells induces gamma rhythm and controls sensory responses. Nature 459, 663-667. Chialvo, D.R. (2010). Emergent complex neural dynamics. Nature Physics 6, 744-750. Colgin, L.L., and Moser, E.I. (2010). Gamma oscillations in the hippocampus. Physiology 25, 319-329. Contreras, D., Destexhe, A., Sejnowski, T.J., and Steriade, M. (1997). Spatiotemporal patterns of spindle oscillations in cortex and thalamus. J Neurosci 17, 1179-1196. Crunelli, V., and Hughes, S.W. (2010). The slow (< 1 Hz) rhythm of non-REM sleep: a dialogue between three cardinal oscillators. Nature neuroscience 13, 9-17. Cunningham, M.O., Whittington, M.A., Bibbig, A., Roopun, A., Lebeau, F.E., Vogt, A., Monyer, H., Buhl, E.H., and Traub, R.D. (2004). A role for fast rhythmic bursting neurons in cortical gamma oscillations in vitro. Proceedings of the National Academy of Sciences of the United States of America 101, 7152-7157. Dugladze, T., Schmitz, D., Whittington, M.A., Vida, I., and Gloveli, T. (2012). Segregation of axonal and somatic activity during fast network oscillations. Science 336, 1458-1461. Fries, P. (2009). Neuronal gamma-band synchronization as a fundamental process in cortical computation. Annual review of neuroscience 32, 209-224. Giraud, A.-L., and Poeppel, D. (2012). Cortical oscillations and speech processing: emerging computational principles and operations. Nature neuroscience 15, 511-517. Goldental, A., Guberman, S., Vardi, R., and Kanter, I. (2014). A computational paradigm for dynamic logic- gates in neuronal activity. Frontiers in computational neuroscience 8. Gray, C.M. (1994). Synchronous oscillations in neuronal systems: mechanisms and functions. J Comput Neurosci 1, 11-38. Grillner, S., Markram, H., De Schutter, E., Silberberg, G., and Lebeau, F.E. (2005). Microcircuits in action– from CPGs to neocortex. Trends in neurosciences 28, 525-533. Hasselmo, M.E. (2005). What is the function of hippocampal theta rhythm?—Linking behavioral data to phasic properties of field potential and unit recording data. Hippocampus 15, 936-949. Jirsa, V.K., and Haken, H. (1996). Field theory of electromagnetic brain activity. Physical Review Letters 77, 960. Kahana, M.J. (2006). The cognitive correlates of human brain oscillations. J Neurosci 26, 1669-1672. Kanter, I., Kopelowitz, E., Vardi, R., Zigzag, M., Kinzel, W., Abeles, M., and Cohen, D. (2011). Nonlocal mechanism for cluster synchronization in neural circuits. EPL (Europhysics Letters) 93, 66001. 20 Cooperative oscillations emerge from response failures Klimesch, W. (1996). Memory processes, brain oscillations and EEG synchronization. International Journal of Psychophysiology 24, 61-100. Klimesch, W. (1999). EEG alpha and theta oscillations reflect cognitive and memory performance: a review and analysis. Brain research reviews 29, 169-195. Marmari, H., Vardi, R., and Kanter, I. (2014). Chaotic and non-chaotic phases in experimental responses of a single neuron. EPL (Europhysics Letters) 106, 46002. Minlebaev, M., Colonnese, M., Tsintsadze, T., Sirota, A., and Khazipov, R. (2011). Early gamma oscillations synchronize developing thalamus and cortex. Science 334, 226-229. Nir, Y., Mukamel, R., Dinstein, I., Privman, E., Harel, M., Fisch, L., Gelbard-Sagiv, H., Kipervasser, S., Andelman, F., and Neufeld, M.Y. (2008). Interhemispheric correlations of slow spontaneous neuronal fluctuations revealed in human sensory cortex. Nature neuroscience 11, 1100-1108. Roxin, A., Riecke, H., and Solla, S.A. (2004). Self-sustained activity in a small-world network of excitable neurons. Physical Review Letters 92, 198101. Sanchez-Vives, M.V., and Mccormick, D.A. (2000). Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nature neuroscience 3, 1027-1034. Silva, L.R., Amitai, Y., and Connors, B.W. (1991). Intrinsic oscillations of neocortex generated by layer 5 pyramidal neurons. Science 251, 432-435. Thivierge, J.-P., Comas, R., and Longtin, A. (2014). Attractor dynamics in local neuronal networks. Frontiers in neural circuits 8. Vardi, R., Goldental, A., Guberman, S., Kalmanovich, A., Marmari, H., and Kanter, I. (2013a). Sudden synchrony leaps accompanied by frequency multiplications in neuronal activity. Frontiers in neural circuits 7. Vardi, R., Goldental, A., Marmari, H., Brama, H., Stern, E.A., Sardi, S., Sabo, P., and Kanter, I. (2015). Neuronal Response Impedance Mechanism Implementing Cooperative Networks with Low Firing Rates and Microsecond Precision. Frontiers in Neural Circuits 9, 29. Vardi, R., Guberman, S., Goldental, A., and Kanter, I. (2013b). An experimental evidence-based computational paradigm for new logic-gates in neuronal activity. EPL (Europhysics Letters) 103, 66001. Vardi, R., Timor, R., Marom, S., Abeles, M., and Kanter, I. (2012a). Synchronization with mismatched synaptic delays: A unique role of elastic neuronal latency. EPL (Europhysics Letters) 100, 48003. Vardi, R., Wallach, A., Kopelowitz, E., Abeles, M., Marom, S., and Kanter, I. (2012b). Synthetic reverberating activity patterns embedded in networks of cortical neurons. EPL (Europhysics Letters) 97, 66002. Wang, X.-J. (2010). Neurophysiological and computational principles of cortical rhythms in cognition. Physiological reviews 90, 1195-1268. Wiest, M.C., and Nicolelis, M.A. (2003). Behavioral detection of tactile stimuli during 7–12 Hz cortical oscillations in awake rats. Nature neuroscience 6, 913-914. Wilson, H.R., and Cowan, J.D. (1972). Excitatory and inhibitory interactions in localized populations of model neurons. Biophysical journal 12, 1. Wu, J.-Y., Guan, L., and Tsau, Y. (1999). Propagating activation during oscillations and evoked responses in neocortical slices. The Journal of neuroscience 19, 5005-5015. 21
1904.02523
1
1904
2019-04-04T12:53:35
Explaining versus Describing Human Decisions. Hilbert Space Structures in Decision Theory
[ "q-bio.NC", "cs.AI", "quant-ph" ]
Despite the impressive success of quantum structures to model long-standing human judgement and decision puzzles, the {\it quantum cognition research programme} still faces challenges about its explanatory power. Indeed, quantum models introduce new parameters, which may fit empirical data without necessarily explaining them. Also, one wonders whether more general non-classical structures are better equipped to model cognitive phenomena. In this paper, we provide a {\it realistic-operational foundation of decision processes} using a known decision-making puzzle, the {\it Ellsberg paradox}, as a case study. Then, we elaborate a novel representation of the Ellsberg decision situation applying standard quantum correspondence rules which map realistic-operational entities into quantum mathematical terms. This result opens the way towards an independent, foundational rather than phenomenological, motivation for a general use of quantum Hilbert space structures in human cognition.
q-bio.NC
q-bio
Explaining versus Describing Human Decisions Hilbert Space Structures in Decision Theory Sandro Sozzo School of Business and Centre IQSCS University Road LE1 7RH Leicester (United Kingdom) Email address: [email protected] Abstract Despite the impressive success of quantum structures to model long-standing human judgement and decision puzzles, the quantum cognition research programme still faces challenges about its explanatory power. Indeed, quantum models introduce new parameters, which may fit empirical data without nec- essarily explaining them. Also, one wonders whether more general non-classical structures are better equipped to model cognitive phenomena. In this paper, we provide a realistic-operational foundation of decision processes using a known decision-making puzzle, the Ellsberg paradox, as a case study. Then, we elaborate a novel representation of the Ellsberg decision situation applying standard quantum corre- spondence rules which map realistic-operational entities into quantum mathematical terms. This result opens the way towards an independent, foundational rather than phenomenological, motivation for a general use of quantum Hilbert space structures in human cognition. Keywords: Quantum structures; Cognitive science; Decision theory; Ellsberg paradox; Operational real- ism. 1 Introduction Traditional cognitive theories systematically apply classical set-theoretic structures to model human judge- ments and decisions under uncertainty. This is particularly evident in theories of rational decision-making, like expected utility theory, where Bayesian, or Kolmogorovian [1], models of probability directly follow from axioms on agents' preferences [2, 3]. However, several cognitive puzzles have been discovered in empirical tests, which provide evidence of systematic deviations from Kolmogorovian probability structures (see, e.g., [4]). For example, Kahneman and Tversky identified a conjunction fallacy in human probability judgements, namely, the law of mono- tonicity of Kolmogorovian probability does not generally hold in this kind of judgements [5]. Also, in human decision-making, Tversky and Shafir proved that the law of total Kolmogorovian probability does not hold in the disjunction effect [6], while Allais and Ellsberg indicated that people do not always choose by maximizing an expected utility with respect to a Kolmogorovian probability measure [7]. As a consequence of the puzzles above, traditional theories using Kolmogorovian structures, though normatively compelling, are descriptively flawed, which led several authors to elaborate alternative pro- posals able to more efficiently and realistically represent human behaviour. This was the starting point of the bounded rationality research programme, initially proposed by Herbert Simon [8] and systematically ap- plied by Kahneman and Tversky [5, 6] to describe concrete judgements and decisions. Bounded rationality models give good predictions in a variety of circumstances. However, despite their simplicity and intuitive 1 character, these models lack a unitary methodology, as well as deeper explanations, and thus provide a very fragmented picture of cognitive phenomena [9]. The quantum cognition research programme has recently attracted the interest of the scientific commu- nity due the superiority of quantum models over traditional and bounded rationality models to deal with the puzzles above. Quantum models were successfully applied to a variety of complex cognitive processes, including human language [10, 11, 12, 13], judgement [4, 14, 15] and decision [15, 16, 17] (see also [18]). Despite these impressive results, however, quantum cognition still raises doubts regarding its explanatory power. Indeed, on the one side, quantum cognitive models introduce new parameters, which may fit ex- perimental data, but do not necessarily explain them. On the other side, one is naturally led to wonder whether cognitive science really needs the entire mathematical formalism of quantum theory in Hilbert space or, on the contrary, non-Kolmogorovian non-Hilbertian models of probability are needed (see, e.g., [19, 20]). In the present paper, we present binding motivations towards an independent, foundational, rather than purely phenomenological, justification of the quantum formalism in human judgement and decision- making under uncertainty. We start from the realistic and operational axiomatizations of quantum physics initiated by Jauch [21] and Piron [22] in Geneva and extended by Aerts (see, e.g., [23, 24]) in Brussels. Efforts have been made in the second part of the last century to derive the mathematical formalism of quantum theory in Hilbert space from more intuitive and empirically justified axioms, resting on basic notions directly connected with the operations that are performed in a laboratory. Particularly, in the Brussels approach, any physical entity is expressed in terms of the basic notions of state, context, property and mutual relationships between them (SCoP system). The approach is realistic, in the sense that the state, being the result of an effective preparation procedure, describes aspects of the reality of the entity. The approach is also operational, in the sense that all basic notions are expressed in terms of well defined empirical terms, like preparation and registration devices, statistics of outcomes, etc. If suitable "purely operational" axioms are imposed on a SCoP system, then the Hilbert space representation uniquely arises for the physical entity. We believe that the above realistic-operational justification of the quantum Hilbert space formalism in physics also provides a strong motivation, if not a justification in itself, for the use of quantum Hilbert space structures in cognition. To this end we particularize in Sect. 4 to a specific decision-making situation, the Ellsberg paradox situation, used as a case study here, the realistic-operational foundation of cognitive entities we have recently elaborated [25], in which a cognitive entity is abstractly described in terms of well defined empirical notions, i.e. state, context, property and outcome probability. Then, the stunning analogies in the realistic and operational descriptions of entities in physical and cognitive realms, suggest that the same Hilbert space leading axiomatics should be used for a cognitive, e.g., decision-making, entity (Sect. 2). The Ellsberg paradox is reviewed in Sect. 3, where we explain the difficulties of both expected utility and bounded rationality theories, to accommodate Ellsberg preferences and the results of more general Ellsberg-like decision situations. We then elaborate in Sect. 5 a mathematical representation in Hilbert space of the Ellsberg paradox situation and the ambiguity aversion pattern found in empirical literature. We had already presented quantum models of various Ellsberg thought experiments, including two-color and three-color urns [16, 17, 26, 27, 28]. The novelty of the mathematical representation developed here consists in the fact that it follows directly from the canonical quantum representation of the realistic-operational terms of state, context, property and outcome probability in Hilbert space, which makes the use of quantum mathematics in this kind of situations more firmly founded and generalizable to other decision situations. We finally offer some conclusive remarks and considerations in Sect. 6, where we specify that the realistic-operational foundation of cognitive science can be in principle extended to several other judgement 2 and decision-making situations, which constitutes a strong indication that "possible failures of Hilbert space modelling" should be searched in other cognitive domains than individual judgements and decisions. 2 Descriptive versus explanatory power of quantum structures Traditional theories of individual judgement and decision-making use, often implicitly, set-theoretic struc- tures, that closely resemble the formal operations of classical Boolean logic and Kolmogorovian probability theory [1]. This is specially evident in rational decision theory, according to which rational agents behave in such a way to maximize expected utility with respect to a Kolmogorovian probability measure and an underlying economic model [2, 3]. These theories are normatively compelling, however, the judgement and decision puzzles in Sect. 1 make them descriptively problematical and suggest alternative more realistic approaches to human behaviour under uncertainty. A major research programme of this kind was initiated by Herbert Simon who put forward the bounded rationality project [8]. Boundedly rational agents experience practical limitations in formulating and solving complex problems and in processing information. They tackle such limitations by taking mental short-cuts, making subjective evaluations and putting psychological aspects above rational reasoning. Within the bounded rationality project, one can cope with cognitive puzzles with judgement heuristics and reasoning biases, namely, the conjunction fallacy with the representativeness heuristics [5], the Allais paradox with prospect theory [6], the Ellsberg paradox with cumulative prospect theory [6], the disjunction effect by uncertainty aversion [6], etc. These approaches undoubtedly provide an intuitive account of how individuals actually behave in situations of uncertainty. However, the reader recognizes at once that a rather eclectic methodology or, better, a variety of methodologies, are employed to accommodate the puzzles above and, while some authors support the hypothesis of an adaptive toolbox to deal with these problems [29], many psychologists will find the bounded rationality research programme as unsatisfactory, while many philosophers of science will try to derive these puzzling phenomena from a universal theory able to overcome the fragmentation of existing approaches. The quantum cognition research programme reaches both effectiveness and unitarity. Since the nineties, quantum Hilbert space models have shown impressive superiority over traditional and bounded rationality approaches in dealing with the puzzles of human cognition and attributing them to genuine quantum effects, like contextuality, emergence, entanglement, interference and superposition. On the other side, quantum models introduce new parameters which can be possibly fitted by empirical data, without however nec- essarily explaining them. Hence, the quantum cognition research programme, though phenomenologically successful, does not seem to offer a deeper understanding and/or explanation of these puzzles. In addi- tion, it is reasonable to wonder whether one really needs the entire Hilbert space formalism to represent cognitive phenomena, and should not better use more general non-Kolmogorovian representations outside physics (see, e.g., [19, 20]). In this respect, it should be noted that prospect theory already proposes non-Kolmogorovian probability models of probability of human decision [6]. It is clear from the considerations above that one needs a deeper justification for the use of the Hilbert space formalism of quantum theory in cognition and decision and, more important, of its necessity. In this respect, a crucial result comes from increasing evidence that "judgements and decisions create rather than record" [4] -- see. e.g., the following quotations. "There is a growing body of evidence that supports an alternative conception according to which preferences are often constructed not merely revealed in the elicitation process. These constructions are contingent on the framing of the problem, the method of elicitation, and the context of the choice." [30] 3 ". . . the process of choice -- and in particular the act of choice -- can make substantial difference to what is chosen. . . . , there is a particular necessity to take note of (i) chooser dependence, and (ii) menu dependence, of preference, even judged from a particular person's perspective." [31] ". . . valuations are initially malleable but become imprinted after the agent is called upon to make an initial decision." [32] It is more and more acknowledged that, in any judgement or decision, a contextual interaction occurs between the situation that is the object of the evaluation and the individual who takes the decision (agent, decision-maker), which may affect the situation itself. At the end of this interaction, a result is actualized among a set of results that were only potential before the interaction [26]. Hence, a judgement/decision process closely resembles a quantum measurement process, where a contextual interaction occurs between the quantum particle that is measured and the measurement apparatus, which changes the state of the quantum particle determining what Heisenberg called "transition from potential to actual". We believe that these analogies between micro-physics and cognition are a good starting point towards a foundational justification for the use of Hilbert space quantum formalism in cognition and decision. In the sixties and seventies of the previous century, several authors wondered whether and how one can provide an independent justification for the Hilbert space formalism in quantum physics, deriving this for- malism from physically justified axioms, resting on well defined empirical notions, directly connected with the operations that are usually performed in a laboratory. One of the well-known approaches to the foun- dations of quantum physics is the Geneva-Brussels realistic-operational approach, initiated by Jauch [21] and Piron [22] in Geneva, and successively extended by Aerts in Brussels (see, e.g., [23, 24]). This research consisted in abstractly describing any physical entity by relevant sets of states, contexts, properties and statistical connections between these notions (SCoP system). These theoretical notions are directly inter- pretable on physical operations on macroscopic apparatuses, such as preparation and registration devices, performed in spatio-temporal domains, such as physical laboratories. Measurements, state transformations, outcome probabilities and dynamics can then be expressed in terms of these more fundamental notions. If suitable axioms are imposed on the mathematical structures underlying a SCoP system, then the Hilbert space structure of quantum theory emerges as a unique mathematical representation, up to isomorphisms [33]. This justification provides the "fundamental architecture of quantum theory in Hilbert space". We have recently proved that any cognitive entity Ω, e.g., a concept, a conceptual combination, a proposition, or a more complex decision-making situation, can be abstractly described by a SCoP system (Σ, L , C , µ, ν) [25], exactly like in physics. We review the essential elements of a SCoP system in cognition in the following. (1) The complex of experimental procedures conceived by the experimenter, the experimental setting and the cognitive effect that one wants to analyse, define a cognitive entity Ω, and are usually associated with a preparation procedure of a state of Ω. (2) Σ is the set of all states of Ω. A state p of Ω is the consequence of a preparation procedure of Ω and has a cognitive, rather than physical, nature. The state of the cognitive entity is a state of affairs. It indeed expresses a "reality of the cognitive entity", in the sense that, once prepared in a given state, such condition is independent of any measurement procedure, and can be confronted with the different participants in an experiment, leading to outcome data and their statistics. (3) C is the set of all contexts of Ω. A context e is an element that can provoke a change of state of the cognitive entity. A special context is the one introduced by a measurement. Indeed, when the cognitive experiment starts, an interaction occurs between the measured entity Ω under study and a participant in the experiment, in which the state p of Ω generally changes, being transformed to another state q. This cognitive interaction is formalized by means of a context e. 4 (4) L is the set of all properties of Ω. A property a of Ω is something Ω "has" independently of any context influencing the entity. An entity Ω is a given state p has a set of properties that are actual in that state, the others being potential. A context e may change the status actual/potential of a property, but cannot change the property itself. (5) The change function µ : Σ × C × Σ −→ [0, 1] is such that, for every p, q ∈ Σ, e ∈ C , µ(q, e, p) is the probability, as the large number limit of relative frequencies, that the context e changes the initial state p of Ω to the final state q. Once recognizes at once in (1) -- (5) the building blocks of the realistic-operational description of a physical entity, in the sense that in both physical and cognitive realms, a SCoP system incorporates all what is needed to study what an entity is, behaves and changes under a context. These impressive analogies indicate that the axioms generally used to justify the Hilbert space formalism of quantum physics are also appropriate to represent cognitive entities and processes. This provides an independent foundational clue and non-phenomenological motivation, if not a justification, that the mathematics of Hilbert space should be used to represent judgement and decision phenomena. In Sect. 4, we will provide a realistic-operational description of a specific decision-making situation, the Ellsberg paradox, setting the grounds for a quantum mathematical representation of it in Sect. 5. In the next section, we will instead summarize the serious difficulties of both traditional and bounded rationality approaches to handle such kind of decision-making situations. 3 Rational decision theory and its puzzles Traditional theories of rational decision-making rest on the tenet that, in situations of uncertainty, indi- vidual agents choose in such a way to maximize their expected utility, or degree of satisfaction. In 1944, von Neumann and Morgenstern presented in a seminal work the first axiomatic formulation of expected utility theory. People continuously take decisions among different options. These decisions are assumed to reveal underlying preferences. Then, von Neumann and Morgenstern proposed a set of "reasonable" axioms on human preferences such that, if the decisions are coherent, in the sense that they reveal axiom satisfying preferences, then the decisions are equivalent to the maximization of an expected utility functional with respect to a Kolmogorovian probability measure [2]. von Neumann and Morgenstern's formulation of expected utility theory has a major limitation, in that it only deals with the uncertainty that can be formalized by known probabilities (also referred to as objective uncertainty, or risk). On the other hand, situations frequently occur in which uncertainty cannot be formalized by known probabilities (also referred to as subjective uncertainty, or ambiguity) [34]. The Bayesian approach to probability minimizes the distinction between objective and subjective uncertainty introducing the notion of subjective probability. Even when probabilities are not known, people may still construct their own beliefs, or priors (which may differ from one individual to another), and they maximize expected utility with respect to these priors. Indeed, Leonard Savage presented in 1954 an axiomatic formulation of expected utility theory which extends the one of von Neumann and Morgenstern to subjective uncertainty [3]. We summarize in the following the essential definitions and results of Savage's expected utility theory, together with its major pitfalls. We refer to [35, 36] for detailed reviews of these results. Savage introduced a set of basic notions, including states of nature, consequences, preferences, and looked for justified axioms on preferences able to provide a representation theorem in which ordering of preferences is characterized by maximization of expected utility. This procedure formally resembles the procedures used in the axiomatizations of quantum physics in Sect. 4. Let S be the set of all (physical) states of nature, which we assume to be discrete and finite here, for the sake of simplicity. Let P(S ) be the power set of S and A ⊆ P(S ) be a (Boolean) σ-algebra. An 5 element E ∈ A denotes an event. A Kolmogorovian probability measure over A is a function p : A ⊆ P(S ) −→ [0, 1] satisfying the axioms of Kolmogorov [1]. Then, let X be the set of all consequences, whose elements we assume to denote monetary payoffs, hence real numbers, here, for the sake of simplicity. In Savage's formulation, a function f : S −→ X denotes an act. Let F be the set of all acts. Let us endow F with a weak preference relation (cid:37), that is, a reflexive, symmetric and transitive relation over the Cartesian product F × F . In (cid:37), the relations (cid:31) and ∼ denote strong preference and indifference, respectively, that is, we write f (cid:31) g whenever an individual strictly prefers act f to act g and f ∼ g whenever the individual is indifferent between f and g. Next, let u : X −→ (cid:60) be a utility function over X . This function typically expresses the decision- maker's taste, hence it is assumed to be strictly increasing and continuous, with additional technical constraints related to the specification of the decision-maker's attitude towards risk. The mathematical definitions above can be simplified by introducing a set {E1, . . . , En} of mutually exclusive and exhaustive elementary events, where Ei = {si ∈ S }, i ∈ {1, . . . , n}, which thus form a partition of S . For every i ∈ {1, . . . , n}, let xi be the utility associated by the act f to the event Ei. Then, f can be equivalently expressed by the 2n-tuple f = (E1, x1; . . . ; En, xn), meaning that the individual will get the outcome x1 if the event E1 occurs (i.e. the state of nature s1 realizes), . . . , the outcome xn if the event En occurs (i.e. the state of nature sn realizes). Finally, we denote by W (f ) = (cid:80)n i=1 p(Ei)u(xi) the expected utility associated with the act f with respect to the Kolmogorovian probability measure p. In his representation theorem, Savage proved that, if the algebraic structure (F , (cid:37)) satisfies a number of "reasonable" axioms1 then, for every f, g ∈ F , a unique Kolmogorovian probability measure p and a unique (up to positive affine transformations) utility function u exist such that f is preferred to g, i.e. f (cid:37) g, if and only if the expected utility of f is greater than the expected utility of g, i.e. W (f ) ≥ W (g). For every i ∈ {1, . . . , n}, the utility value u(xi) depends on the decision-maker's risk preferences, while p(Ei) is interpreted as the subjective probability, expressing the individual's belief that the event Ei occurs [3]. Savage's result is both compelling at a normative level and testable at a descriptive level. Indeed: (i) if the axioms are intuitively reasonable and decision-makers agree with them, then they must all behave as if they were maximizing an expected utility with respect to a single subjective probability distribution satisfying Kolmogorov's axioms; (ii) the axioms suggest to design decision-making experiments to test the validity of expected utility theory, hence of the axioms themselves, in real life situations. Because of (i), Savage's expected utility formulation is generally accepted to prescribe "how rational agents should choose". However, one the one side, the theory offers very little about where beliefs come from and how they should be calculated and, on the other side, regarding (ii), decision-making experiments have systematically found deviations from that rational behaviour in concrete situations. In particular, Daniel Ellsberg proved in 1961 in a number of thought experiments that decision-makers generally prefer acts with known (or objective) probabilities over acts with unknown (or subjective) prob- abilities [7]. We analyse here the famous Ellsberg three-color example as a paradigmatic example to show that (i) traditional decision theories do not work, (ii) bounded rationality approaches are not sufficiently explanatory, (iii) quantum structures are needed. Consider one urn with 30 red balls and 60 balls that are either yellow or black, the latter in unknown proportion. One ball will be drawn at random from the urn. Then, free of charge, a person is asked to bet 1One of the axioms is the famous sure thing principle, which is violated in the Ellsberg paradox. The other axioms are: ordinal event independence, comparative probability, non-degeneracy, small event continuity and dominance, and have a technical nature [3]. However, these axioms are not relevant to the present purposes, hence we will not dwell on them, for the sake of brevity. 6 Act f1 f2 f3 f4 1/3 2/3 Red $100 $0 $100 $0 Yellow $0 $0 $100 $100 Black $0 $100 $0 $100 Table 1. The payoff matrix for the Ellsberg three-color example. on pairs of the acts f1, f2, f3 and f4 in Table 1. Ellsberg suggested that, when asked to rank these acts, most individuals will prefer f1 over f2 and f4 over f3. Indeed, f1 and f4 are unambiguous acts, in the sense that they are associated with events over known probabilities -- the events "a red ball is drawn" and "a yellow or black ball is drawn" are associated with objective probabilities 1/3 and 2/3, respectively. On the contrary, f2 and f3 are ambiguous acts, in the sense that they are associated with events over unknown probabilities -- the events "a yellow ball is drawn" and "a black ball is drawn" are both associated with a probability ranging from 0 to 2/3. This attitude of decision-makers to prefer "probabilized over non-probabilized uncertainty" has been known as ambiguity aversion since Ellsberg studies [7]. Several experiments on Ellsberg urns decisions, but also on financial, insurance and medical decisions, have confirmed the Ellsberg preferences f1 (cid:31) f2 and f4 (cid:31) f3, thus indicating that ambiguity aversion is a good candidate to explain concrete decisions in this case, and only Slovic and Tversky found ambiguity seeking patterns (see, e.g., [36] for a review of experimental studies). In [27], we tested various human decision puzzles, including the Ellsberg three-color example. We asked 200 people, chosen among colleagues and friends, to fill a questionnaire in which they had to choose between various options. People had on average a basic knowledge of probability theory, but no specific training in decision theory. Participants were provided with a questionnaire similar to the one in Figure 1, in which they had to choose between acts f1 and f2 and, then, between acts f3 and f4 in Table 1. Overall, 125 participants preferred acts f1 and f4, 38 preferred acts f1 and f3, 6 preferred acts f2 and f3, and 31 preferred acts f2 and f4. This means that 163 participants over 200 preferred act f1 over act f2, which entails a preference weight of 0.815. Also, 156 participants over 200 preferred act f4 over act f3, which entails a preference weight of 0.780. The inversion rate is 0.655, a pattern that agrees with the Ellsberg preferences found in the literature and significantly indicates the presence of ambiguity aversion. Preferences of decision-makers who are sensitive to ambiguity, that is, are ambiguity averse or ambiguity seeking, cannot be explained within Savage's expected utility theory, because they violate the sure thing principle, according to which, preferences should be independent of the common outcome. In the specific case of the three-color example, preferences should not depend on whether the common event "a yellow ball is drawn" pays off $0 or $100. More technically, Savage's expected utility theory predicts consistency of preferences, namely, f1 is preferred to f2 if and only if f3 is preferred to f4. A simple calculation shows that this is impossible within a traditional expected utility framework. Indeed, if we denote by pR, pY and pB the probability that a red ball, a yellow ball, a black ball, respectively, is drawn (with pR = 1/3 = 1 − (pY + pB)), then the expected utilities W (fi), i = 1, 2, 3, 4, are such that W (f1) > W (f2) if and only if (pR − pB)(u(100)− u(0)) > 0 if and only if W (f3) > W (f4). We can equivalently say that no assignment of Kolmogorovian probabilities pR, pY and pB reproduces a preference with W (f1) > W (f2) and W (f4) > W (f3), whence the Ellsberg paradox. Several extensions of Savage's expected utility theory have been put forward in order to accommodate the Ellsberg paradox (see, e.g., the reviews in [35] and [36]). One of the major proposals is Tversky and Kahneman's cumulative prospect theory, mentioned in Sect. 2 and elaborated within a bounded rationality research programme [6]. In particular, to reproduce Ellsberg preferences, Tversky and Kahneman replaced 7 Figure 1. A sample of the questionnaire related to the decision-making experiment on the Ellsberg three-color example: choice between acts f1 and f2 in Table 1. 8 (i) the utility function u by a scaling function u(cid:48) reflecting the subjective value of the outcome utility, and (ii) the subjective probability measure p by a non-additive measure p(cid:48) satisfying the mathematical properties of a capacity. As we have mentioned in Sect. 2, such bounded rationality models, though descriptively interesting and easily interpretable intuitively, provide a too fragmented view of decision theory, hence they are not able to provide a unitary and adequate explanatory framework to understand the deep aspects of decision processes. In addition, cumulative prospect theory, as well as other major non-expected utility models, fails to reproduce the empirical results of a recently elaborated variant of the Ellsberg paradox, the Machina paradox [27, 37, 38]. An innovative aspect of descriptive, like bounded rationality, approaches, is the representation of subjec- tive probabilities by more general, possibly non-Kolmogorovian, mathematical structures. This is crucial towards a more satisfactory framework for human decision-making that goes beyond Savage's expected utility, as we will see in Sect. 5. In the next section, we intend to elaborate a realistic-operational description of a decision-making situation, using the Ellsberg three-color example as a case study. We will demonstrate that, once the Ellsberg paradox situation is formulated in terms of states, contexts, properties and transition probabilities, then the application of the mathematical formalism of quantum theory directly follows from the canonical representation of these realistic-operational terms in Hilbert space. 4 A realistic-operational description of a decision-making situation In this section, we specify the realistic-operational description of cognitive entities in Sect. 2 to the decision- making situation presented in the Ellsberg three-color example [25]. In it, we explicitly distinguish physical from cognitive, in this case, decision-making, entities. Analogously, we distinguish physical from cognitive states of nature, though one can intuitively see that some cognitive states are mapped into the corresponding physical states. In the Ellsberg three-color example, the cognitive, i.e. decision-making, entity ΩDM is the urn with 30 red balls and 60 yellow or black balls in unknown proportion. This is what the individual reads in a questionnaire, interacts with and takes a decision on. The cognitive entity ΩDM is associated with a defined set ΣDM of states.2 A state p ∈ ΣDM has a cognitive nature and incorporates aspects of ambiguity. A context e does not pertain to ΩDM but can interact with it. Let CDM be the set of all contexts of ΩDM . The interaction of ΩDM with a context e ∈ CDM may determine a change of the state of ΩDM from p to a different state q. The probability of such a state transition will be denoted by µ(q, e, p), where µ : ΣDM ×CDM ×ΣDM −→ [0, 1]. We might complete the realistic-operational description of ΩDM defining a set LDM of properties and an actuality relation connecting properties and states. However, they are not needed in the Ellsberg three-color scenario, hence we omit specifying these notions here, for the sake of brevity, though they may be relevant in more general decision situations. Let us now introduce a color context eC ∈ CDM , which is the context associated with a drawing of a ball from the urn. As a result of the drawing, we have three possible outcomes, R, Y and B, corresponding to the colors of the balls, red, yellow and black, respectively. The outcomes R, Y and B are the eigenvalues of eC and are respectively associated with the final states, or eigenstates, pR, pY and pB of the cognitive entity ΩDM . These eigenstates are such that µ(pi, eC, pi) = 1, i ∈ {R, Y, B}. 2Some authors identify the notion of "state" with the notion of "belief state" of the individual participating in the cognitive experiment, e.g., taking the decision (see, e.g., [4, 9, 15]). We instead neatly distinguish states from measurements here. A state is defined by a preparation procedure of the cognitive entity under investigation. The participant in the experiment acts as a (measurement) context that interacts with the cognitive entity and changes its state. 9 The color context eC introduces three mutually exclusive and exhaustive elementary events Ei = (eC,{i}), i ∈ {R, Y, B}, which are such that the subjective probability that the event Ei occurs when the cognitive entity ΩDM is in the state p is given by the transition probability µ(pi, eC, p) ∈ [0, 1], i ∈ {R, Y, B}. Then, in analogy with Savage's expected utility theory, we can introduce monetary payoffs x ∈ X ⊆ (cid:60), utility functions u : X −→ (cid:60) and acts taking the form f = (ER, xR; EY , xY , EB, xB), mapping the event Ei into the payoff xi, i ∈ {R, Y, B}. In particular, the acts f1, f2, f3 and f4 in Table 1, Sect. 3 are defined in the way above. Let us now come to the operational-realistic description of a decision-making process in the Ellsberg three-color situation. Suppose that, in the absence of any context, the entity ΩDM is in the initial state p0. This state corresponds to a preparation of the cognitive entity ΩDM and can be set by the information on the corresponding physical entity. For example, it is reasonable to assume that p0 is such that, for every i ∈ {R, Y, B}, µ(pi, eC, p0) = 1/3, because of the indifference principle. Whenever an individual is asked to rank f1 and f2, the individual's attitude towards ambiguity, e.g., ambiguity aversion, can be described as a new context e1 ∈ CDM acting on ΩDM in the initial state p0 and changing p0 into a new state p1, characterized by a new probability distribution µ(pi, eC, p1), i ∈ {R, Y, B}. Similarly, whenever the individual is asked to rank f3 and f4, the individual's attitude towards ambiguity, e.g., ambiguity aversion, can be described as a new context e2 ∈ CDM acting on ΩDM in the initial state p0 and changing p0 into a new state p2, characterized by a new probability distribution µ(pi, eC, p2), i ∈ {R, Y, B}. The cognitive states p1 and p2, and their ensuing probability distributions, are responsible of the inversion of preferences which occur in the Ellsberg paradox situation. Finally, a decision process between acts f1 and f2 can be operationally described as a measurement context e12 ∈ CDM acting on the entity ΩDM in the ambiguity averse state p1, with possible outcomes "yes" and "no". Similarly, a decision process between acts f3 and f4 can be described as a measurement context e34 ∈ CDM acting on the entity ΩDM in the ambiguity averse state p2, with possible outcomes "yes" and "no". These contexts give rise of the statistics of outcomes in a decision-making test on the Ellsberg three-color urn. Now, the considerations in Sect. 2 naturally indicate to represent states, contexts, properties and outcome probabilities of ΩDM by using the canonical quantum representation of states, contexts, properties and outcome probabilities in Hilbert space. In particular, subjective probabilities will be represented using the Born rule of quantum probability. This is what we intend to show in the next section where the realistic- operational terms defined here will be canonically represented using quantum mathematical terms. 5 A novel quantum representation of the Ellsberg paradox In this section we elaborate a new quantum representation of the three-color example straightly follow- ing the canonical Hilbert space representation of the realistic and operational notions in Sect. 3. This representation generalizes and strengthens those in [16, 17, 26, 27]. 5.1 Quantum representation of basic notions The cognitive entity ΩDM is associated with a Hilbert space H . Since, the three-color example involves three mutually exclusive and exhaustive elementary events, H can be chosen to be isomorphic to the complex Hilbert space C3 of ordered triples of complex numbers. Let {(1, 0, 0), (0, 1, 0), (0, 0, 1)} be the canonical orthonormal basis of C3. A state pv of the entity ΩDM is represented by the unit vector v(cid:105) ∈ H , v(cid:105) =(cid:112)(cid:104)vv(cid:105) = 1. The elementary event Ei is represented by the one-dimensional orthogonal projection operator Pi = i(cid:105)(cid:104)i, i ∈ {R, Y, B}, where we choose R(cid:105) = (1, 0, 0), Y (cid:105) = (0, 1, 0) and B(cid:105) = (0, 0, 1). The color context 10 eC is then represented by the spectral family {PR, PY , PB}. In the canonical basis of C3, the unit vector v(cid:105) can be written as v(cid:105) = ρReiθRR(cid:105) + ρY eiθY Y (cid:105) + ρBeiθBB(cid:105) = (ρReiθR, ρY eiθY , ρBeiθB ) (1) We use the Born rule to represent subjective probabilities. Then, the subjective probability that the elementary event Ei, i ∈ {R, Y, B}, occurs when the entity ΩDM is in the state pv is given by µv(Ei) = (cid:104)vPiv(cid:105) = (cid:104)iv(cid:105)2 = ρ2 i (2) In addition, the subjective probability that the event E, represented by the orthogonal projection operator PE, occurs when the cognitive entity ΩDM is in the state pv is µv(E) = (cid:104)vPEv(cid:105) = PEv(cid:105)2. Finally, for every state pv represented by the unit vector v(cid:105), the subjective probability measure (3) is a quantum probability measure over the lattice L (C3) of all orthogonal projection operators on the Hilbert space C3. Compatibility with the standard three-color situation entails ρ2 R = 1 3 , hence it follows from Eq. (1) that µv : L (C3) −→ [0, 1] v(cid:105) = ( where the last equality is obtained from(cid:112)(cid:104)vv(cid:105) = 1. eiθR, ρY eiθY , ρBeiθB ) = ( 1√ 3 1√ 3 eiθR, ρY eiθY , Special states are the state with no black balls represented by and the state with no yellow balls represented by vRY (cid:105) = ( 1√ 3 eiθR, (cid:114) 2 3 eiθY , 0) (cid:114) 2 (cid:114) 2 3 − ρ2 yeiθB ) (4) (5) (6) (7) (8) (9) The acts f1, f2, f3 and f4 are represented by the self-adjoint operators vRB(cid:105) = ( 1√ 3 eiθR, 0, eiθB ) 3 F1 = u(100)PR + u(0)PY + u(0)PB F2 = u(0)PR + u(0)PY + u(100)PB F3 = u(100)PR + u(100)PY + u(0)PB F4 = u(0)PR + u(100)PY + u(100)PB (10) respectively. The utility function u(·) is not given by the theory but it is revealed in a decision test by concrete choices, in analogy with standard procedures in the literature. The expected utility Wv(fi), i ∈ {1, 2, 3, 4}, in a generic state pv of the entity ΩDM is Wv(f1) = (cid:104)v F1v(cid:105) = 1 3 Wv(f2) = (cid:104)v F2v(cid:105) = ( 1 3 Wv(f3) = (cid:104)v F3v(cid:105) = ( 1 3 Wv(f4) = (cid:104)v F4v(cid:105) = 1 3 u(100) + 2 3 u(0) + ρ2 Y )u(0) + ( 2 3 + ρ2 Y )u(100) + ( − ρ2 2 3 Y )u(100) − ρ2 Y )u(0) u(0) + 2 3 u(100) 11 (11) (12) (13) (14) As we can see, the expected utilities Wv(f1) and Wv(f4) do not depend on the cognitive state pv of the entity ΩDM , in agreement with the fact that f1 and f4 are unambiguous acts. On the contrary, the expected utilities Wv(f2) and Wv(f3) do depend on the cognitive state pv, in agreement with the fact that f2 and f3 are ambiguous acts. This also agrees with our assumption that cognitive states provide information on ambiguity. 5.2 Reproducing Ellsberg preferences with ambiguity averse states Let us now suppose that, in the absence of any context, the cognitive entity ΩDM is in the initial state p0. The principle of indifference (see Sect. 4) then suggests that p0 is the state represented by the unit vector v0(cid:105) = 1√ 3 (1, 1, 1) (15) leading to uniform probabilities of drawing a red, yellow and black ball. The ambiguity attitude contexts e1 and e2 will determine a change of state of the entity ΩDM , depending on individual preferences to- wards ambiguity. For example, two ambiguity seeking states pw1 and pw2 will be such that the following inequalities hold Ww1(f2) > Ww2(f3) > 1 3 1 3 u(100) + 2 3 u(0) u(0) + 2 3 u(100) (16) (17) We will instead explicitly determine two ambiguity averse states pw1 and pw2 which reproduce Ellsberg preferences, that is, Two general cognitive states pw1 and pw2 states are represented by the unit vectors Ww1(f1) > Ww1(f2) Ww2(f4) > Ww2(f3) w1(cid:105) = ( w2(cid:105) = ( 1√ 3 1√ 3 eiθR, ρY eiθY , eiφR, τY eiφY , − ρ2 yeiθB ) − τ 2 y eiφB ) (cid:114) 2 (cid:114) 2 3 3 (18) (19) (20) (21) (22) (23) respectively. For the sake of simplicity, let us look for states with simple phases, namely, θR = φR = 0 and θY , θB, φY , φB = 0, π. In particular, one can show that, for every α > 1√ , the unit vectors 3 (cid:114) 2 (cid:114) 2 3 − α2) 3 − α2, α) w1(cid:105) = ( w2(cid:105) = ( 1√ 3 1√ 3 , α,− ,− reproduce Ellsberg preferences. However, the ambiguity averse states pw1 and pw2 are not generally or- thogonal, unless α = ±0.7887. Let us choose the positive sign, so that the orthonormal vectors w1(cid:105) = (0.5774, 0.7887,−0.2113) w2(cid:105) = (0.5774,−0.2113, 0.7887) (24) (25) reproduce Ellsberg preferences in the three-color example within a quantum mathematical representation. 12 5.3 Modelling empirical data in Hilbert space The final step of the quantum representation of the three-color example consists in modelling the experimen- tal data in Sect. 3. To this aim, we describe the decision between acts f1 and f2 by a measurement context e12 and represent the latter by the spectral family {M, 1 − M}, where the orthogonal projection operator M projects onto the one-dimensional subspace generated by the unit vector m(cid:105) = ( 1√ , ρY eiθY , ρBeiθB ), where ρ2 3 Y + ρ2 B = 2 3 . The one-dimensional projection operator M is then given by ρBe−iθB 1√ 3 ρY ρBei(θY −θB) M = m(cid:105)(cid:104)m = ρY e−iθY ρ2 Y 1√ 3 1 3 1√ 3 1√ 3 ρY eiθY ρBeiθB ρY ρBe−i(θY −θB) ρ2 B Analogously, we describe the decision between acts f3 and f4 by a measurement context e34 and represent the latter by the spectral family {N, 1 − N}, where the orthogonal projection operator N projects onto the one-dimensional subspace generated by the unit vector n(cid:105) = ( 1√ B = 2 3 . The one-dimensional projection operator N is then given by , τY eiφY , τBeiφB ), where τ 2 Y + τ 2 3     N = n(cid:105)(cid:104)n = 1 3 1√ 3 τY e−iφY τ 2 Y τBe−iφB 1√ 3 τY τBei(φY −φB) τY eiφY τBeiφB 1√ 3 1√ 3 (1, 1, 1) and w1(cid:105) and w2(cid:105) in Eqs. (24) and (25). It follows that the τY τBe−i(φY −φB) (27) τ 2 B We refer to the unit vectors v0(cid:105) = 1√ conditions 3 (cid:104)mm(cid:105) = 1 (cid:104)nn(cid:105) = 1 (cid:104)w1Mw1(cid:105) = 0.815 (cid:104)w2Nw2(cid:105) = 0.780 (cid:104)v0Mv0(cid:105) = 0.500 (cid:104)v0Nv0(cid:105) = 0.500 (28) (29) (30) (31) (32) (33) must be satisfied by the real parameters ρY , ρB, θY , θB, τY , τB, φY and φB. Equations (28) and (29) are determined by normalization conditions, while Eqs. (30) and (31) are determined by empirical data. Finally, Eqs. (32) and (33) are determined by the fact that decision-makers who are not sensitive to ambiguity should overall be indifferent between f1 and f2, as well as between f3 and f4. Hence, on average, half respondents are expected to prefer f1 (f3) and the other half f2 (f4). To simplify the analysis, let us set θY = φB = 0. Hence, we are left with a system of 6 equations in 6 unknown variables whose solution is (26) (34) (35) (36)  ρY = 0.6853 ρB = 0.4438 θB = 105.07◦ τY = 0.4785 τB = 0.6616 φY = 102.87◦ Equivalently, we get m(cid:105) = (0.5773, 0.6853, 0.4438ei105.07◦ n(cid:105) = (0.5773, 0.4785ei102.87◦ , 0.6616) ) 13 Thus, the orthogonal projection operators in Eqs. (26) and (27) reproducing the experimental data in Sect. 3 are 0.333 0.396 0.256ei105.07◦ 0.333 0.276ei102.87◦ 0.382 0.396 0.470 0.304ei105.07◦ 0.276e−i102.87◦ 0.317e−i102.87◦ 0.229 0.256e−i105.07◦ 0.304e−i105.07◦ 0.197 0.382 0.317ei102.87◦ 0.438     M = N = (37) (38) The construction of a quantum model for the data on the Ellsberg three-color experiment in Sect. 3 is thus completed. As we can see, the quantum model naturally arises from the canonical Hilbert space representation of the realistic-operational termsin Sect. 4. 6 Conclusions Despite its phenomenological success to deal with classically problematical cognitive puzzles, the quan- tum cognition research programme still poses challenging questions regarding its explanatory power and necessity. In this paper, we specialized to the Ellsberg paradox decision situation a realistic-operational foundation which we have recently extended from physics to cognition. Then, we applied to the Ellsberg three-color example the canonical quantum representation of realistic-operational terms in Hilbert space. This result on the Ellsberg paradox situation is paradigmatic, in the sense that one can follow the same strategy to generally claim that the mathematical representation of human judgements and decision-making in Hilbert space has now an independent motivation of a foundational, rather than phenomenological, nature. To conclude, we agree that quantum theory in Hilbert space is not the ultimate theory in cognition -- recent results on sequential measurements and order effects seem to confirm this conclusion(see, e.g., [25]). However, we also believe that there are strong theoretical motivations, in addition to its empirical success and unitary explanation, to continue using Hilbert space structures in cognition. Acknowledgements This work was supported by QUARTZ (Quantum Information Access and Retrieval Theory), the Marie Sk(cid:32)lodowska-Curie Innovative Training Network 721321 of the European Union's Horizon 2020 research and innovation programme. References [1] Kolmogorov, A.N.: Grundbegriffe der Wahrscheinlichkeitrechnung, Ergebnisse Der Mathematik (1933); translated as: Foundations of Probability. Chelsea Publishing Company, New York (1950) [2] von Neumann, J., Morgenstern, O.: Theory of Games and Economic Behavior. Princeton University Press, Princeton (1944) [3] Savage, L.: The Foundations of Statistics. John Wiley & Sons, New York (1954); revised and enlarged edition: Dover Publications, New York (1972) 14 [4] Busemeyer, J.R., Bruza, P.D.: Quantum Models of Cognition and Decision. Cambridge University Press, Cambridge (2012) [5] Kahneman, D., Slovic, P., Tversky, A.: Judgment Under Uncertainty: Heuristics and Biases. Cam- bridge University Press, New York (1982) [6] Kahneman, D., Tversky, A. (Eds.): Choices, Values and Frames. Cambridge University Press, New York (2000) [7] Ellsberg, D.: Risk, ambiguity, and the Savage axioms. Q. J. Econ. 75, 643 -- 669 (1961) [8] Simon, H: A behavioral model of rational choice. Quart. J. Econ. 69, 99-118 (1955) [9] Blutner, R., beim Graben, P.: Quantum cognition and bounded rationality. Synthese 193, 3239 -- 3291 (2016) [10] Dalla Chiara, M.L., Giuntini, R., Leporini, R., Sergioli, G.: Holistic logical arguments in quantum computation. Mathematica Slovaca 66, 313 -- 334 (2006) [11] Aerts, D.: Quantum structure in cognition. J. Math. Psychol. 53, 314 -- 348 (2009) [12] Aerts, D., Gabora, L., Sozzo, S.: Concepts and their dynamics: A quantum -- theoretic modeling of human thought. Top. Cogn. Sci. 5, 737 -- 772 (2013b) [13] Dalla Chiara, M.L., Giuntini, R., Leporini, R., Negri, E., Sergioli, G.: Quantum information, cognition, and music. Front. Psychol. doi 10.3389/fpsyg.2015.01583 (2015) [14] Aerts, D., Broekaert, J., Gabora, L., Sozzo, S.: Quantum structure and human thought. Behav. Brain Sci. 36, 274 -- 276 (2013a) [15] Haven, E., Khrennikov, A.Y.: Quantum Social Science. Cambridge University Press, Cambridge (2013) [16] Aerts, D., Sozzo, S., Tapia, J.: Identifying quantum structures in the Ellsberg paradox. Int. J. Theor. Phys. 53, 3666 -- 3682 (2014) [17] Aerts, D., Sozzo, S.: From ambiguity aversion to a generalized expected utility. Modeling preferences in a quantum probabilistic framework. J. Math. Psychol. 74, 117 -- 127 (2016) [18] Sozzo, S.: Effectiveness of the quantum-mechanical formalism in cognitive modeling. Soft Comp. 21, 1455 -- 1465 (2017) [19] Holik, F., Fortin, S., Bosyk, G., Plastino, A.: On the interpretation of probabilities in generalized probabilistic models. Lect. Not. Comp. Sci. 10106, 194 -- 205 (2016) [20] Holik, F., Sergioli, G., Freytes, H., Plastino, A.: Pattern recognition in non-Kolmogorovian structures. Found. Sci. 23, 119 -- 132 (2017) [21] Jauch, J.M.: Foundations of Quantum Mechanics. Addison Wesley, Reading MA (1968) [22] Piron, C.: Foundations of Quantum Physics. Benjamin, Reading MA (1976) [23] Aerts, D.: Foundations of quantum physics: A general realistic and operational approach. Int. J. Theor. Phys. 38, 289 -- 358 (1999) 15 [24] Aerts, D.: Being and change: Foundations of a realistic operational formalism. In: Probing the Structure of Quantum Mechanics: Nonlinearity, Nonlocality, Computation and Axiomatics. Aerts, D., Czachor, M., Durt, T. (Eds.) World Scientific, Singapore (2002) [25] Aerts, D., Sassoli de Bianchi, M., Sozzo, S.: On the foundations of the Brussels operational-realistic approach to cognition. Front. Phys. 4, doi: 10.3389/fphy.2016.00017 (2016) [26] Aerts, D., Haven, E., Sozzo, S.: A proposal to extend expected utility in a quantum probabilistic framework. Econ. Theory 65, 1079 -- 1109 (2018) [27] Aerts, D., Geriente, S., Moreira, C., Sozzo, S.: Testing ambiguity and Machina preferences within a quantum-theoretic framework for decision-making, J. Math. Econ. doi 10.1016/j.jmateco.2017.12.002 (2018) [28] Sozzo, S.: Quantum structures in human decision-making: Towards quantum expected utility, Int. J. Theor. Phys. doi 10.1007/s10773-019-04022-w (2019) [29] Gigerenzer, G., Selten, R.: Bounded Rationality: An Adaptive Toolbox. MIT Press, Cambridge MA (2001) [30] Tversky, A., Simonson, I.: Context-dependent preferences. Manag. Sci. 39, 85 -- 117 (1993) [31] Sen, A.: Maximization and the act of choice. Econometrica 65, 745 -- 779 (1997) [32] Ariely, D., Prelec, G., Lowenstein, D.: "Coherent arbitrariness": Stable demand curve without stable preferences. Quart. J. Econ. 118, 73-103 (2003) [33] Beltrametti, E.G., Cassinelli, G.: The Logic of Quantum Mechanics. Addison -- Wesley, Reading MA (1981) [34] Knight, F.H.: Risk, Uncertainty and Profit. Houghton Mifflin, Boston (1921) [35] Gilboa, I., Marinacci, M.: Ambiguity and the Bayesian paradigm. In: Advances in Economics and Econometrics: Theory and Applications. Acemoglu, D., Arellano, M., Dekel, E. (Eds.) Cambridge University Press, New York (2013) [36] Machina, M.J., Siniscalchi, M.: Ambiguity and ambiguity aversion. In: Machina, M.J., Viscusi, K. (Eds.) Handbook of the Economics of Risk and Uncertainty, pp. 729 -- 807. Elsevier, New York (2014) [37] Machina, M.J.: Risk, ambiguity, and the dark -- dependence axioms. Am. Econ. Rev. 99, 385-392 (2009) [38] L'Haridon, O., Placido, L.: Betting on Machinas reflection example: An experiment on ambiguity. Theor. Decis. 69, 375 -- 393 (2010) 16
1712.10062
1
1712
2017-12-28T21:26:43
Multi-timescale memory dynamics in a reinforcement learning network with attention-gated memory
[ "q-bio.NC", "cs.LG", "cs.NE", "stat.ML" ]
Learning and memory are intertwined in our brain and their relationship is at the core of several recent neural network models. In particular, the Attention-Gated MEmory Tagging model (AuGMEnT) is a reinforcement learning network with an emphasis on biological plausibility of memory dynamics and learning. We find that the AuGMEnT network does not solve some hierarchical tasks, where higher-level stimuli have to be maintained over a long time, while lower-level stimuli need to be remembered and forgotten over a shorter timescale. To overcome this limitation, we introduce hybrid AuGMEnT, with leaky or short-timescale and non-leaky or long-timescale units in memory, that allow to exchange lower-level information while maintaining higher-level one, thus solving both hierarchical and distractor tasks.
q-bio.NC
q-bio
Multi-timescale memory dynamics in a reinforcement learning network with attention-gated memory Marco Martinolli†, Wulfram Gerstner† and Aditya Gilra† †School of Computer and Communication Sciences, and Brain-Mind Institute, School of Life Sciences, École Polytechnique Fédérale de Lausanne, 1015 Lausanne EPFL, Switzerland Correspondence: [email protected], [email protected] Abstract Learning and memory are intertwined in our brain and their relationship is at the core of several recent neural network models. In particular, the Attention-Gated MEmory Tagging model (AuGMEnT) is a reinforcement learning network with an emphasis on biological plausibility of memory dynamics and learning. We find that the AuGMEnT network does not solve some hierarchical tasks, where higher-level stimuli have to be maintained over a long time, while lower-level stimuli need to be remembered and forgotten over a shorter timescale. To overcome this limitation, we introduce hybrid AuGMEnT, with leaky or short-timescale and non-leaky or long-timescale units in memory, that allow to exchange lower-level information while maintaining higher-level one, thus solving both hierarchical and distractor tasks. 1 Introduction Memory spans various timescales and plays a crucial role in human and animal learning [Tetzlaff et al. 2012]. In cognitive neuroscience, the memory system that enables manipulation and storage of information over a period of a few seconds is called Working Memory (WM), and is correlated with activity in prefrontal cortex (PFC) and basal ganglia (BG) [Frank et al. 2001; Mink 1996]. In com- putational neuroscience, there are not only several standalone models of WM dynamics [Barak and Tsodyks 2014; Samsonovich and McNaughton 1997; Compte et al. 2000], but also supervised and reinforcement learning models augmented by working memory [Alexander and Brown 2015; Rombouts et al. 2015; Graves et al. 2014; 2016; Santoro et al. 2016]. Memory mechanisms can be implemented by enriching a subset of artificial neurons with slow time constants and gating mechanisms [Hochreiter and Schmidhuber 1997; Gers and Schmidhuber 2001; Cho 2014]. More recent memory-augmented neural network models like Neural Turing Ma- chine [Graves et al. 2014] and Differentiable Neural Computer [Graves et al. 2016], employ an address- able memory matrix that works as a repository of past experiences and a neural controller that is able to store and retrieve information from the external memory to improve its learning performance. Here, we study and extend the Attention-Gated MEmory Tagging model or AuGMEnT [Rombouts et al. 2015]. AuGMEnT is trained with a Reinforce- ment Learning (RL) scheme, where learning is based on a reward signal that is released after each response selection. The representation of stimuli is accumulated in the memory states and the mem- ory is reset at the end of each trial (see Methods). The main advantage of the AuGMEnT network for the computational neuroscience community resides in the biological plausibility of its learning algorithm. Notably, the AuGMEnT network [Rombouts et al. 1 2015] uses a memory-augmented version of a biologically plausible learning rule [Roelfsema and van Ooyen 2005] mimicking backpropagation (BP). Learning is the result of the joint action of two factors, neuromodulation and attentional feedback, both influencing synaptic plasticity. The former is a global reward-related signal that is released homogeneously across the network to inform each synapse of the reward prediction error after response selection [Schultz et al. 1993; 1997; Waelti et al. 2001]. Neuromodulators such as dopamine influence synaptic plasticity [Yagishita et al. 2014; He et al. 2015; Brzosko et al. 2015; 2017; Frémaux and Gerstner 2016]. The novelty of AuGMEnT compared to three-factor rules [Xie and Seung 2004; Legenstein et al. 2008; Vasilaki et al. 2009; Frémaux and Gerstner 2016] is to add an attentional feedback system in order to keep track of the synaptic connections that cooperated for the selection of the winning action and overcome the so-called structural credit assignment problem [Roelfsema and van Ooyen 2005; Rombouts et al. 2015]. AuGMEnT includes a memory system, where units accumulate activity across several stimuli in order to solve temporal credit assignment tasks involving delayed reward delivery [Sutton 1984; Okano et al. 2000]. The attentional feedback mechanism in AuGMEnT works with: a) synaptic eligibility traces that decay slowly over time, and b) non-decaying neuronal traces that store the history of stimuli presented to the network up to the current time [Rombouts et al. 2015] [Rombouts et al. 2015]. The AuGMEnT network solves the Saccade-AntiSaccade task [Rombouts et al. 2015], which is equivalent to a temporal XOR task [Ab- bott et al. 2016] (see Supplementary Material). different decay constants so that they work on different temporal scales, while the network learns to weight their usage based on the requirements of the specific task. In our simulations, we employed just two subgroups of cells in the memory, where one half of the memory is non-leaky and the other is leaky with a uniform decay time constant; however, more generally, the hybrid AuGMEnT architecture may contain several subgroups with distinct leakage behaviours. The paper is structured as follows. Section 2 presents the architectural and mathematical details of hybrid AuGMEnT. Section 3 describes the simulated results of the hybrid AuGMEnT network, the standard AuGMEnT network and a fully leaky control network, on two cognitive tasks, a non- hierarchical task involving sequence prediction [Cui et al. 2015] and a hierarchical task 12AX [O'Reilly and Frank 2006]. Finally, in Section 4 we discuss our main achievements in comparison with state-of-the-art models and present possible future developments of the work. 2 Methods 2.1 Hybrid AuGMEnT network architecture and operation The network controls an agent which, in each time step t, receives a reward in response to the previ- ous action, processes the next stimulus, and takes the next action, as in Figure 1B. In each time step, we distinguish two phases, called the feedforward pass and feedback pass, depicted in Figure 1C. However, in the case of more complex tasks with long trials and multiple stimuli, like 12AX [O'Reilly and Frank 2006] depicted in Figure 1A, we find that the accumulation of information in AuGMEnT can lead to memory interference and loss in performance. Hence, we ask the question whether a modified AuGMEnT model would lead to a broader applicability of attention-gated rein- forcement learning. We propose a variant of the AuGMEnT network, named hybrid AuGMEnT, that introduces timescales of forgetting or leakage in the memory dynamics to overcome this kind of learning limitation. We employ memory units with 2.1.1 Feedforward pass: stimulus to action selec- tion In AuGMEnT [Rombouts et al. 2015], information is processed through a network with three layers, as shown in the left panel of Figure 1C. Each unit of the output layer corresponds to an action. There are two pathways into the output layer: the regular R branch and the memory M branch. The regular branch is a standard feedforward network with one hidden layer. The current stimu- lus sR i (t), indexed by unit index i = 1, . . . , S is con- 2 B A C Figure 1. Overview of AuGMEnT network operation. A. Example of trials in the 12AX task, where task symbols appear sequentially on a screen organized in outer loops, which start with either digit 1 or 2, and a random number of inner loops (e.g. B-Y, C-X and A-X). Each cue presentation is associated with a Target (R) or Non-Target (L) correct response. When output and correct response coincide, the agent receives a positive reward (+), otherwise it gets a negative reward (-). Figure is adapted from Figure 1 of O'Reilly and Frank [2006]. B. AuGMEnT operates in discrete time steps each comprising the reception of reward (r), input of state or stimulus (s) and action taken (a). It implements the State-Action-Reward-State-Action (SARSA, in figure s'a'rsa) reinforcement learning algorithm. In time step t, reward r is obtained for the previous action a' taken in time step t − 1. The network weights are updated once the next action a is chosen. C. The AuGMEnT network is structured in three layers with different types of units. Each iteration of the learning process consists of a feedforward pass (left) and a feedback pass (right). In the feedforward pass, sensory information about the current stimulus in the bottom layer, is fed to regular units without memory (left branch) and units with memory (right branch) in the middle layer, whose activities in turn are weighted to compute the Q-values in the top activity layer. Based on the Q-values, the current action is selected (e.g. green z2). The reward obtained for the previous action is used to compute the reward prediction error, which modifies the connection weights, that contributed to the selection of the previous action, in proportion to their eligibility traces. After this, temporal eligibility traces and tags (in green) on the connections are updated to reflect the correlations between the current pre and post activities. Then, in the feedback pass, spatial eligibility traces (in red) are updated, attention-gated by the current action (e.g. red z2), via feedback weights. 3 nected to the hidden units (called regular units) in- dexed by j, via a set of modifiable synaptic weights ji yielding activity yR vR j : j (t) = σ(cid:0)hR j (cid:1) , yR (cid:88) hR j = vR jisR i (t), (1) i where σ is the sigmoidal function σ(x) = (1 + exp(−x))−1. Input units are one-hot bi- nary with values Si ∈ {0, 1} (equal to 1 if stimulus i is currently presented, 0 otherwise). The memory branch is driven by transitions be- tween stimuli, instead of the stimuli themselves. The sensory input of the memory branch consists of a set of 2S transient units, i.e. S ON units l ∈ {0, 1}S that encode the onset of each stimu- s+ lus, and S OFF units s− l ∈ {0, 1}S that encode the offset: l (t) = [sl(t) − sl(t − 1)]+ s+ s− l (t) = [sl(t − 1) − sl(t)]+, (2) where the brackets signify rectification. In the following, we denote the input into the memory branch with a variable sM i defined as the concate- nation of these ON and OFF units: sM i (t) = s+ i (t), s− i−S(t), if i ≤ S if i > S, (3) (cid:40) (cid:88) The memory units have to maintain task-relevant information through time. The transient input is transmitted via the synaptic connections vM ji to the memory layer, where it is accumulated in the states: hM j (t) = ϕjhM j (t − 1) + vM ji sM i (t). (4) i We introduce the factor ϕj ∈ [0, 1] here, as an extension to the standard AuGMEnT [Rombouts et al. 2015], to incorporate decay or forgetting of j over time. Setting ϕj ≡ 1 the memory state hM for all j, we obtain non-leaky memory dynamics as in the original AuGMEnT network [Rombouts et al. 2015] (Fig. In our hybrid AuGMEnT network, each memory cell or subgroup of memory cells may be assigned different leak co-efficients ϕj (Fig. 2, right panel). In this way, the memory is composed of subpopulations of neurons that cooperate in different ways to solve 2, left panel). 4 a task, allowing at the same time long-time main- tenance and fast decay of information in memory. In contrast to the forget gate of Long Short-Term Memory [Hochreiter and Schmidhuber 1997] or Gated Recurrent Unit [Cho 2014], our memory leak co-efficient is not trained and gated, but fixed. The memory state hM j leads to the activation of a memory unit: j (t) = σ(cid:0)hM j (t)(cid:1) . yM (5) The states of the memory units are reset to 0 at the end of each trial. Both branches converge onto the output layer. The activity of an output unit with index k approx- imates the Q-value of action k = a given the input s ≡ [si], denoted as Qs,a(t). Q-values are formally defined as the future expected discounted reward conditioned on stimulus s(t) and action a(t), that is: (cid:12)(cid:12)(cid:12)(cid:12) s = s(t), a = a(t) (cid:35) (6) Qs,a(t) = E γτ rt+τ +1 (cid:34) ∞(cid:88) τ =0 where γ ∈ [0, 1] is a discount factor. Numeri- cally, the vector Q that approximates the Q-values is obtained by combining linearly the hidden states from the regular and the memory branches, with synaptic weights wR kj and wM kj : Qk(t) = wR kjyR j (t) + wM kj yM j (t). (7) j j Finally, the Q-values of the different actions partic- ipate in an -greedy winner-take-all competition to select the response of the network. With probabil- ity 1 − , the next action a(t) is the one with the maximal Q-value: a(t) = argmaxkQk(t). (8) With probability , a stochastic policy is chosen with probability: (cid:88) (cid:88) (cid:80) pa = exp(g(t)Qa) k exp(g(t)Qk) (9) π arctan( t is a weight function defined as where g(t) g(t) = 1 + 10 t∗ ), that gradually increases in time over a task-specific, fixed time scale t∗. Over time, this emphasizes the action with maximal Q-value, improving prediction stability. AuGMEnT Hybrid AuGMEnT Figure 2. Architectures of standard AuGMEnT and hybrid AuGMEnT networks. The difference between the net- works consist in their memory dynamics: the memory layer of standard AuGMEnT (left) has only conservative units with ϕj ≡ 1, while hybrid AuGMEnT (right) possesses a memory composed of both leaky ϕj < 1 and non-leaky ϕj = 1 units. 2.1.2 Feedforward pass: reward-based update of weights, and correlation-based update of eligibility traces and tags AuGMEnT follows the SARSA updating scheme and updates the Q-values for the previous action a(cid:48) taken at time t − 1, once the action a at time t is known (see Fig. 1B). Q-values depend on the weights via equation (7). The temporal difference (TD) error is defined as [Wiering and Schmidhuber 1998; Sutton and Barto 1998]: δ(t) = (r(t) + γQa(t)) − Qa(cid:48)(t − 1), (10) where a is the action chosen at current time t, and r(t) is the reward obtained for the action a(cid:48) taken at time t − 1. The Temporal Difference (TD) error δ(t) acts as a global reinforcement signal to modify the weights of all connections as vR,M ji wR,M kj (t + 1) = vR,M (t + 1) = wR,M ji kj (t) + βeR,M (t) + βeR,M ji kj (t)δ(t), (t)δ(t), (11) and eR,M where β is a learning rate and eR,M are synaptic eligibility traces, defined below. Superscript R or M denotes the regular or memory branch respectively. We use the same symbol eR,M for eligibility traces at the input-to-hidden (i to kj ji 5 j) and hidden-to-output (j to k) synapses, even though these are different, with the appropriate one clear from context and the convention for indices. After the update of weights, a synapse from neu- ron j in the hidden layer to neuron k in the output layer updates its temporal eligibility trace eR kj(t + 1) = yR kj (t + 1) = yM eM j (t)zk(t) + (1 − α)eR j (t)zk(t) + (1 − α)eM kj(t), kj (t), (12) where α ∈ [0, 1] is a decay parameter, zk is a binary one-hot variable that indicates the winning action (equal to 1 if action k has been selected, 0 otherwise), and M or R denotes the regular or the memory branch respectively. Similarly, a synapse from neuron i in the input layer to neuron j in the hidden layer sets momen- tary tags T R,M as: ji j (t)), i (t) σ(cid:48)(hR ji (t) σ(cid:48)(hM ji (t) = sR T R ji (t) = X M T M where σ(cid:48)(hR,M ) is a nonlinear function of the in- put potential, defined as the derivative of the gain j (t)), (13) j function σ, and X M is a synaptic trace [Pfister and ji Gerstner 2006; Morrison et al. 2008] defined as fol- lows: i j ji X M (14) i (t). , zk, sR,M ji (t) = ϕj X M ji (t − 1) + sM Note that the tag T R,M has no memory beyond one time step, i.e. it is set anew at each time step. Nev- ertheless, since X M ji depends on previous times, the tag T M ji of memory units can link across time steps. Since activities yR,M and input potentials hR,M are quantities available at the synapse, a bi- j ological synapse can implement the updates of eli- gibility traces and tags locally. We emphasize that both eligibility traces and tags can be interpreted as 'Hebbian' correlation detectors. In the original AuGMEnT model [Rombouts et al. 2015], all eligibil- ity traces and tags were said to be updated in the feedback pass. Here, without changing the order of operations of the algorithm, we have conceptu- ally shifted the update of those traces and tags that depend on the correlations of the activities, to the last step of the feedforward pass when these activ- ities are still available. Note that activities could in principle change via attention-gating during the feedback pass [Roelfsema et al. 2010; Moore and Armstrong 2003]. jk It must be noted that the feedback synapses w(cid:48)R,M follow the same update rule as their feed- forward partner wR,M Therefore, even if the initializations of the feedforward and feedback weights are different, their strengths become similar during learning, as suggested by neuro- physiological findings [Mao et al. 2011]. kj . 2.2 Deriving the learning rules via gradient descent For networks with one hidden layer and one- hot coding in the output, attentional feedback is equivalent to backpropagation [Roelfsema and van Ooyen 2005; Rombouts et al. 2015]. We can show that the equations for eligibility traces and tagging along with the weight update equations reduce an RPE-based loss function E defined as: E = 1 2 (δ(t))2 (16) Here we specifically discuss the case of the tagging equations (13) and (15) and the update rule (11) associated with the weight vM ji from sensory input into memory, as it contains the memory decay factor ϕj that we introduced, but analogous dis- cussion holds also for weights vR kj and wR kj. ji, wM 2.1.3 Feedback pass: attention-gated update of eligibility traces Proof. We want to show that After action selection and the updates of weights, tags, and temporal eligibility traces in the feedfor- ward pass, the synapses that contributed to the cur- rently selected action update their spatial eligibil- ity traces in an attentional feedback step. For the synapses from the input to the hidden layer, the tag T R,M from equation (13) is combined with a spa- ji tial eligibility trace which can be interpreted as an attentional feedback signal [Rombouts et al. 2015]. eR ji(t + 1) = T R ji eM ji (t + 1) = T M ji w(cid:48)R jkzk + (1 − α)eR jk zk + (1 − α)eM w(cid:48)M ji(t), ji (t), (cid:88) (cid:88) k k ∆vM ji = β eM ji δt ∝ − ∂E ∂vM ji (17) For simplicity, here we prove (17) neglecting the temporal decay of the eligibility trace eM ji (i.e. α = 1), so that eM ji = T M ji (cid:48)M jk zk = T M ji w (cid:48)M ja(cid:48) w where a(cid:48) is the selected action at time t − 1. k (cid:88) (15) We first observe that the right-hand side of equa- tion (17) can be rewritten as: where feedback weights from the output layer to the hidden layer have been denoted as w(cid:48) jk and zk ∈ {0, 1} is the value of output unit k (one-hot response vector as defined for equation (12)). − ∂E ∂vM ji = − ∂E ∂Qa(cid:48) ∂Qa(cid:48) ∂vM ji = δt ∂Qa(cid:48) ∂vM ji Thus, it remains to show that ∂Qa(cid:48) ∂vM ji = eM ji . 6 Similarly to the approach used in backpropaga- tion, we now apply the chain rule and we focus on each term separately: ∂Qa(cid:48) ∂vM ji = ∂Qa(cid:48) ∂yM j ∂yM j ∂hM j ∂hM j ∂vM ji Parameter β : Learning parameter λ : Eligibility persistence γ : Discount factor α : Eligibility decay rate  : Exploration rate Value 0.15 0.15 0.9 1 − γλ 0.025 From equations (5) and (7), we immediately have that: ∂yM j ∂hM j = σ(cid:48)(hM j ) ∂Qa(cid:48) ∂yM j = wM a(cid:48)j However, in the feedback step the weight wM a(cid:48)j is replaced by its feedback counterpart w(cid:48)M ja(cid:48) . As dis- cussed above, this is a valid approximation because they become similar during learning. Finally, starting from equation (4) we can write: Table 1. Parameters for the AuGMEnT network. memory cells and ϕj = 0.7 for the second half. To reduce to the standard AuGMEnT [Rombouts et al. 2015] network, we set ϕj ≡ 1 for all j, while for a leaky control network we set ϕj ≡ 0.7 for all j. In general, the leak co-efficients may be tuned to adapt the overall memory dynamics to the specific task. (cid:88) ≈(cid:88) i t−1(cid:88) (cid:88) t(cid:88) τ =t0 i ϕt−τ j sM i (τ ) hM j (t) = vM ji (t) sM i (t) + vM ji (t) i τ =t0 ϕt−τ j vM ji (τ ) sM i (τ ) 3 Results where t0 indicates the starting time of the trial and last approximation derives from the assumption of slow learning dynamics, i.e. vM ij (t) for t0 ≤ τ < t. As a consequence, we have: ij (τ ) = vM ∂hM ∂vM j (t − 1) ji (t − 1) ≈ t−1(cid:88) τ =t0 ϕt−τ +1 j sM i (τ ) = X M ji (t − 1) In conclusion, we combine the different terms and we obtain the desired result: (cid:48)M ja(cid:48) = δt eM ji . ji ∝ δt X M ji σ(cid:48)(hM j ) w ∆vM 2.3 Simulation and tasks simulation scripts were written (https://www.python.org/), in All with python rendered using the matplotlib module plots (http://matplotlib.org/). simulation and plotting scripts are available online at https://github.com/martin592/hybrid_AuGMEnT. These We use the parameters listed in Table 1 for our simulations. Further, for the Hybrid AuGMEnT network, we set ϕj = 1 for the first half of the 7 AuGMEnT [Rombouts et al. 2015] includes a differ- entiable memory system and is trained in an RL framework with learning rules based on the joint effect of synaptic tagging, attentional feedback and neuromodulation (see Methods). Here, we study our proposed variant of AuGMEnT, named hybrid AuGMEnT, that has an additional leak factor in a subset of memory units, and compare it to the original AuGMEnT and to a control network with all leaky memory units. As a first step, we validated our implementa- tions of standard and hybrid AuGMEnT networks on the Saccade-AntiSaccade (S-AS) task, used in the reference paper [Rombouts et al. 2015] (Supplementary Material). We next simulated the networks on two other cognitive tasks with differ- ent structure and memory demands: the sequence prediction task [Cui et al. 2015] and the 12AX task [O'Reilly and Frank 2006]. In the former, the agent has to predict the final letter of a sequence depend- ing only on its starting letter, while in the latter, the agent has to identify target pairs inside a sequence of hierarchical symbols. The S-AS task maps to a temporal XOR task [Abbott et al. 2016], thus the hidden layer is essential for the task [Minsky and Papert 1969; Rumelhart et al. 1985]. The 12AX also resembles an XOR structure, but is more complex Table 2. Network Architecture Parameters for the Simulations Network Parameter Sequence Prediction Task 12AX Task S : Number of sensory units R : Number of regular units M : Number of memory units A : Number of activity units (L = sequence length) L − 1 3 8 2 8 10 20 2 due to an additional dimension and distractors in the inner loop (Supplementary Figure 3). The complexity of the sequence prediction task is less compared to the 12AX task, and can be effectively solved by AuGMEnT. We will show that hybrid AuGMEnT performs well on both cognitive tasks, whereas standard AuGMEnT fails on the 12AX task. The parameters involving the architecture of the networks on each task are reported in Table 2. We now discuss each of the tasks in more detail. 3.1 Task 1: Sequence Prediction In the sequence prediction task [Cui et al. 2015], letters appear sequentially on a screen and at the end of each trial the agent has to correctly predict the last letter. Each sequence starts either with an A or with an X, which is followed by a fixed sequence of letters (e.g. B-C-D-E). The trial ends with the prediction of the final letter, which depends on the initial cue: if the sequence started with A, then the final letter has to be a Z; if the initial cue was an X, then the final letter has to be a Y. In case of correct prediction the agent receives a reward of 1 unit, otherwise he is punished with a negative reward of −1. A scheme of the task is presented in Figure 3 for sequences of four letters. The network has to learn the task for a given sequence length, kept fixed throughout training. The agent must learn to maintain in memory, the initial cue of the sequence until the end of the trial, to solve the task. At the same time, the agent has to learn to neglect the information coming from the intermediate cues (called distractors). Thus the difficulty of the task is correlated with the length of the sequence. We studied the performance of the AuGMEnT network [Rombouts et al. 2015] and our hybrid Figure 3. Scheme of the sequence prediction task. Scheme of sequence prediction trials with sequence length equal to 4 (i.e. 2 distractors): the two possible sequences are: A-B-C-Z (blue) or X-B-C-Y (red) variant on the sequence prediction task. The mean trend of the RPE-based energy function defined in equation (16) (Fig. 4A) shows that both models converge in a few hundreds of iterations. As a control, we also simulated a variant in which also memory units were leaky. We noticed that hybrid and standard AuGMEnT networks are more efficient than the purely leaky control. This is not surprising because the key point in the sequence prediction task consists in the maintenance of the initial stimulus, which is simpler with a non-leaky memory than with a leaky one. We notice that the hybrid model has a behaviour similar to AuGMEnT. We also analyzed the effect of the temporal length of the sequences on the network perfor- mance, by varying the number of distractors (i.e. the intermediate letters) per sequence (Fig. 4B). For each sequence length, the network was retrained ab initio. We required 100 consecutive correct predictions as the criterion for convergence. We ran 100 simulations starting with different initial- izations for each sequence length and averaged 8 A B Figure 4. Convergence in the sequence prediction task. A. Time course of error of the models on the sequence prediction task with sequences of five letters (three distractors): the mean squared RPE decays to zero for all networks but the leaky control network (blue) is much slower than AuGMEnT (green) and Hybrid AuGMEnT (red). B. Convergence time of the AuGMEnT network and its variants on the sequence prediction task with increasing number of distractors, i.e. intermediate cues before final prediction. Figure 5. Memory weights of AuGMEnT networks in sequence prediction task. Memory weight matrices after convergence for AuGMEnT (left) and Hybrid AuGMEnT (right) networks on the sequence prediction task with se- quences of length five (first row) and ten (second row). Note that the first two memory units in Hybrid AuGMEnT are leaky (M1-L and M2-L), while the last ones are conservative or non-leaky (M1-C and M2-C). 9 Table 3. The 12AX task: table of key information Task feature Input Output Target sequences 1-. . . -A-X or 2-. . . -B-Y. Details 8 possible stimuli: 1,2,A,B,C,X,Y,Z. Non-Target (L) or Target (R). Training dataset Inner loops Probability of target sequence is 25%. Sequence of outer loops starting with 1 or 2. Maximum number of training samples is 1000000. Each outer loop contains a random number of inner loops, between 1 and 4 the convergence time. Again, AuGMEnT and Hy- brid AuGMEnT show good learning performance, maintaining an average of about 250 trials to convergence for sequences containing up to 10 distractors, whereas a network with purely leaky units is much slower to converge. The leaky dynamics are not helpful for the sequence prediction task, because the intermediate cues are not relevant for the final model perfor- mance. Therefore, it is sufficient to supress the weight values in the VM matrix for distractors, and increase those of the initial A/X letter. This is confirmed by the structure of the weight matrices of the memory in all networks shown after con- vergence (Fig. 5) in simulations of the sequence prediction task on sequences of five and ten letters. The weight values are highest in absolute value for letters A and X, for both the ON (+) and OFF (−) units. Finally, we also notice that Hybrid AuGMEnT employs mainly the conservative or non-leaky memory units (M1-C and M2-C) rather than the leaky ones (M1-L and M2-L) to solve the task, showing that the network is able to focus the update dynamics on the connections that are more suitable for the specific task. 3.2 Task 2: 12AX The 12AX task is a standard cognitive task used to test Working Memory and diagnose behavioral and cognitive deficits related to memory dys- functions [Alexander and Brown 2015]. Basically, the problem consists in identifying some target sequences among a group of symbols that appear on a screen. The general procedure of the task is schematized in Figure 1A and details involving the construction of the 12AX dataset are collected in Table 3. The set of possible stimuli consists of 8 symbols: two digit cues (1 and 2), two context cues (A and B), two target cues (X and Y) and finally two distractors (C and Z). Each trial (or outer loop) starts with a digit cue and is followed by a random number Inner loops are composed of 1 to 4 inner loops. of patterns of context-target cues, like A-X, B-X or B-Y. The distractors are non task-relevant cues that can invalidate a subsequence creating wrong inner loops like A-Z or C-X. The cues are presented one by one on a screen and the agent has two possible responses for each of them: Target (R) and Non-Target (L). There are only two valid Target cases: in trials that start with digit 1, the Target is associated with the target cue X if preceded by context A (1-. . . -A-X); otherwise, in case of initial digit 2, the Target occurs if the target cue Y comes after context B (2-. . . -B-Y). The dots are inserted to stress that the target inner loop can occur even a long time after the digit cue, as happens in the following example sequence: 1-A-Z-B-Y-C-X-A-X (whose sequence of correct responses is L-L-L-L- L-L-L-L-R). The variability in the temporal length of each trial is the main issue in solving the 12AX task because of the temporal credit assignment problem. Moreover, since 1-A-X and 2-B-Y are target sequences, whereas 2-A-X and 1-A-Y are not, the task can be seen as a generalization of temporal XOR (Supplementary Fig. 3). The types of the inner loops are determined randomly, with a probability of 50% to have pairs A-X or B-Y. As a result, combined with the 10 probability to have either 1 or 2 as starting digit of the trial, the overall probability to have target pair is 25%. Since the Target response R has to be associated only with an X or Y stimulus that appears in the correct sequence, the number of Non-Targets L is generally much larger, on average 8.96 Non-Targets to 1 Target. We rewarded the correct predictions of a Non-Target with 0.1 and of Targets with 1, and punished the wrong predic- tions with reward of −1. In effect, we balanced the positive reward approximately equally between Targets and Non-Targets based on their relative frequencies, which aids convergence. We simulated Hybrid AuGMEnT network, base AuGMEnT and leaky control on the 12AX task, in order to see whether in this case the introduction of the leaky dynamics improves learning perfor- mance. Figure 6A shows the evolution of the mean squared RPE for the three networks. After a sharp descent, all networks converge to an error level that is non-zero, indicating that learning of the 12AX task is not completely achieved, possibly due to memory interference. However, hybrid AuGMEnT and leaky control saturate at a lower error value than base AuGMEnT. This difference can be attributed mostly to the errors in responding to the Target cues (Fig. 6B) than to Non-Target cues (Fig. 6C). Note that, since 12AX is a Continuous Performance Task (CPT), the error is computed at each iteration – including the more frequent and trivial Non-Target predictions – and averaged over 2, 000 consecutive predictions. All networks quickly learn to recognize the Non-Target cues (1, 2, A, B, C, Z are always Non-Targets) (Fig. 6C). However, hybrid AuGMEnT and leaky control learn the more complex identification of Target patterns within a trial when X or Y are presented to the network, better than base AuGMEnT (Fig. 6B). The gap in the mean squared RPE between hybrid AuGMEnT and leaky control versus base AuGMEnT is wider when only potential target cues are con- sidered in the mean-squared RPE as in Figure 6B, than when only non-targets are considered as in Figure 6C. convergence Adopting the criterion from Alexander and Brown [2015] that requires 1, 000 consecutive correct predictions, we show the per- centage of successful learning over 100 simulations and the average learning time in Figure 7 for the three networks. Standard AuGMEnT network was unable to match the convergence condition during any simulation (0% success), despite presenting 1,000,000 outer loop trials in each simulation. However, hybrid AuGMEnT and leaky control performed 100% consistently, suggesting that leaky memory units are necessary for the 12AX task. Leaky control learned slightly faster (learn- ing time mean=30, 032.2 and s.d.=11, 408.9) than hybrid AuGMEnT (learning time mean=34, 263.6 and s.d.=12, 737.3). 8, left panel): In order to understand how the hybrid memory works on the 12AX task, we analyzed the weight structure of the connectivity matrices which be- long to the memory branch of the hybrid AuGMEnT network (Figure 8). Unlike in the sequence pre- diction task, here the hybrid network employs both the leaky and the non-leaky memory units. However, there is an overall separation in the memory activity between the two groups of cells (Fig. the leaky units are mainly responsible for the storage of the digit information, having the highest values in absolute value on the weights associated with 1(±) and 2(±) (e.g. on M4 and M9), while the non-leaky cells emphasize more the information coming from the potential Target cues X(±) and Y(±) (e.g. on M14, M17 and M20). The storage of the initial digit cue is 'assigned' to leaky units, because the interference from the following letters is reduced thanks to the gradual loss of information while the digit information can survive through sufficient increasing of the digit-related input weights. In this way, the mem- ory interference problem is mitigated, because the crucial digit information is maintained over time without interference in the leaky units and the identification of the inner loops is done by the conservative part of the memory. As a result, all memory units contribute to the definition of the activity Q-values (Fig. 8, right panel) and, in particular, the memory units that are more active (i.e. the same ones mentioned above) are the ones that strongly discriminate Non-Targets (L) against Targets (R), giving positive contribution to one and negative to the other. The memory units show an opposing behavior on activation versus on deactivation of Target cues: 11 A B C Figure 6. Learning convergence of the AuGMEnT variants in the 12AX task. Minimization of the RPE-based energy function during training on the 12AX task. A. All networks show a good decay of the mean-squared RPE, but they seem to converge to a non-zero regime and, in particular, the base AuGMEnT network (green) is the one that maintains a higher mean-squared RPE level when compared to leaky control (blue) and Hybrid AuGMEnT (red). B. Mean-square RPE associated with only potential target cues X and Y. C. Mean-squared RPE related to only non-target cues. Figure 7. Comparative statistics of the AuGMEnT variants on performance on the 12AX task. Barplot description of the learning behavior of the three networks on the 12AX task according to the convergence criterion given by Alexander and Brown [2015]. After 100 simulations, we measured the fraction of times that the model satisfies con- vergence condition (left) and the average number of training trials needed to meet the same convergence criterion (right). Although training dataset consists of 1, 000, 000 outer loops, the base AuGMEnT network never manages to satisfy the convergence criterion, while the leaky (blue) and hybrid (red) models have similar convergence performance with a learning time of about 30, 000 trials. for instance, if X+ has strong positive weight in- tensity, then X- shows a contrary negative weight intensity (see M14 or M17). In this way, the network tries to reduce the problems of memory interfer- ence between subsequent inner cycles by adding to the memory during deactivation, an opposite amount of information stored during the previous activation, effectively erasing the memory. Further, the difference in absolute value between activation and deactivation is higher in case of the leaky cells, because the deactivation at the next iteration has to remove only a lower amount of information from the memory due to leakage. However, for the digit cues 1 and 2, the weights for activation and deactivation have typically the same sign in order to reinforce the digit signal in memory in two subsequent timesteps (see on M4 and M9). 12 Figure 8. Memory weights of Hybrid AuGMEnT in the 12AX task. Plot of the weight matrices in the memory branch of hybrid AuGMEnT network after convergence on the 12AX Task. Left: weights from the transient stimulus into the 20 memory units (half leaky, half conservative). Right: weights from the memory cells into the output units. In conclusion, the conservative dynamics of the memory in standard AuGMEnT can be a limitation for the learning ability of the model, especially in cases of complex tasks with many data to store or long trials. In fact, even though the complexity of the 12AX task is limited compared to other typical RL tasks, the AuGMEnT network fails to maintain a sufficiently stable performance to satisfy the required convergence criterion. The introduction of the leaky co-efficient in Hybrid AuGMEnT leads to the network solving the 12AX task, overcom- ing memory interference. However, the loss of information from the leaky memory does not improve learning in other tasks with lower risk of memory interference like the sequence prediction task. Hybrid AuGMEnTcan be adapted to different task structures and to different temporal scales by varying the size and the composition of the memory, for example by considering multiple subpopulations of neurons with distinct memory timescales, say in a power law distribution. 4 Discussion A key goal of the computational neuroscience community is to develop neural networks that are at the same time biologically plausible and able to learn complex tasks similar to humans. The embedding of memory is certainly an important step in this direction, because memory plays a central role in human learning and decision making. Our interest in the AuGMEnT network [Rombouts et al. 2015] derives mainly from the biological plausibility of its learning and memory dynamics. In particular, the biological setting of the learning algorithm is based on synaptic tag- ging, attentional feedback and neuromodulation, providing a possible biological interpretation to backpropagation-like methods. We developed Hybrid AuGMEnT, by introducing leaky dynamics in the memory system, with the aim of improving its learning performance and extending the variety of solvable tasks. Hybrid AuGMEnT with both leaky and non-leaky units in its memory system, solves the 12AX task 13 on which standard AuGMEnT fails. Both solve the simpler saccade-antisaccade and sequence prediction tasks. Hybrid AuGMEnT inherits the biological plausibility of base AuGMEnT [Rombouts et al. 2015]. In addition, consistent learning with decaying memory units requires that the decay of synaptic traces in a memory unit, as per equation (14) [Pfister and Gerstner 2006; Morrison et al. 2008], be at the same timescale as decay of the unit's memory state as per equation (4). Despite the improvement with our hybrid the learning ability of AuGMEnT is variant, still limited compared to other state-of-the-art memory-augmented networks. For instance, the Hierarchical Temporal Memory (HTM) network [Cui et al. 2015] presents a greater flexibility in sequence learning than what we have experienced in AuGMEnT on the simple sequence prediction task. Utilizing a complex column-based architec- ture and an efficient system of inner inhibitions, the HTM network is able to maintain a dual neural activity, both at column level and at unit level, that allows to have sparse representations of the input and give multi-order predictions using an unsu- pervised Hebbian-like learning rule. Nonetheless, it is unclear how the HTM network can be applied to reward-based learning, in particular to tasks like the 12AX, with variable number of inner loops. Although the hybrid memory in the AuGMEnT network remarkably improved its convergence performance on the 12AX task, its learning effi- ciency is still lower than the reference Hierarchical Error Representation model (HER) [Alexander and Brown 2015; 2016]. In fact, in our simulations, hy- brid AuGMEnT showed a mean time to convergence equal to 34, 263.6 outer loops, while the average learning time of HER on the same convergence condition is around 750 outer loops. The reason for this large gap in the learning performance resides in the gating mechanism of HER network that is specifically developed for hierarchical tasks and is used to decide at each iteration whether to store the new input or maintain the previous content in memory. Unlike HER model, the memory in AuGMEnT does not include any gating mechanism, meaning that the network does not learn when to store and recall information but the memory dy- namics are entirely developed via standard weight modulation. On the other hand, the HER model is not as biologically plausible as the AuGMEnT net- work, because, although its hierarchical structure is inspired on the supposed organization of the prefrontal cortex, its learning scheme is artificial and based on standard backpropagation. In addition, the recent delta-RNN network [Ororbia et al. 2017] presents interesting simi- larities with hybrid AuGMEnT in employing two timescales, maintaining memory via interpolation of fast and slow changing inner representations. The delta-RNN, whose memory dynamics are a generalization of the gating mechanisms of LSTM and GRU, outperforms these popular recurrent architectures. Thus, it likely has a better learning ability than hybrid AuGMEnT, though it requires a higher number of parameters and the network is not based on biological considerations. The lack of a memory gating system is a great limitation for AuGMEnT variants, when compared with networks equipped with a gated memory, like HER [Alexander and Brown 2015; 2016] or LSTM [Hochreiter and Schmidhuber 1997; Gers and Schmidhuber 2001], especially on complex tasks with high memory demand. Still, even though it cannot be properly defined as a gating system, the forgetting dynamics introduced in hybrid AuGMEnT has a similar effect as the activity of the forget gates in LSTM or GRU. However, unlike forget gates, the decay coefficients are not learnable and are not input-dependent for each memory cell. The Hybrid AuGMEnT network could be further developed by adding a gating control on the leakage: leak gates could be an output of the controller branch of the network and then applied as a gate or decay co-efficient in the memory branch. In this way, the gating value becomes stimulus-dependent and leakage is adjusted to optimize the model performance. On the other hand, such a gating system would make the network more complex, where learning of the gate variables implies an error backpropagation through multiple layers, that may compromise the biological plausibility of the AuGMEnT learning dynamics (though see [Lillicrap et al. 2016; Baldi et al. 2016; Guerguiev et al. 2017]). Alternatively, inspired by the hierarchical ar- 14 chitecture of HER [Alexander and Brown 2015], the memory in AuGMEnT could be divided into multiple levels each with their own memory dynamics: each memory level could be associated with distinct synaptic decay and leaky coefficients, learning rates, or gates, in order to cover different temporal scales and encourage level specialization. Compared with hybrid AuGMEnT, this differ- entiation in memory will not only involve the leaky dynamics, but also the temporal dynamics associated with attentional feedback and synaptic potentiation. In the past years, the reinforcement learning community has proposed several deep RL net- works, like deep Q-networks [Mnih et al. 2015] or the AlphaGo model [Chen 2016], that combine the learning advantages of deep neural networks with reinforcement learning [Li 2017]. Thus, it may be interesting to consider a deep version of the AuGMEnT network with additional hidden layers of neurons. While conventional error backpropagation in AuGMEnT may not yield local synaptic plasticity, locality might be retained with alternative backpropagation methods [Lillicrap et al. 2016; Baldi et al. 2016; Guerguiev et al. 2017]. 5 Acknowledgements We thank Vineet Jain for helpful discussions. Financial support was provided by the European Research Council (Multirules, grant agreement no. 268689), the Swiss National Science Foundation (Sinergia, grant agreement no. CRSII2_147636), and the European Commission Horizon 2020 Framework Program (H2020) (Human Brain Project, grant agreement no. 720270). References Abbott, L., De Pasquale, B., and Memmesheimer, R.-M. (2016). Building functional networks of spiking model neurons. Nature neuroscience, 19(3):350–355. Alexander, W. H. and Brown, J. W. (2015). Hier- archical error representation: A computational model of anterior cingulate and dorsolateral pre- frontal cortex. Neural Computation, 27(11):2354– 2410. Alexander, W. H. and Brown, J. W. (2016). Frontal cortex function derives from hierarchical predic- tive coding. bioRxiv, page 076505. Baldi, P., Sadowski, P., and Lu, Z. (2016). Learn- ing in the machine: Random backpropaga- arXiv preprint tion and the learning channel. arXiv:1612.02734. Barak, O. and Tsodyks, M. (2014). Working models of working memory. Current opinion in neurobiol- ogy, 25:20–24. Brzosko, Z., Schultz, W., and Paulsen, O. (2015). Retroactive modulation of spike timing- eLife, page dependent plasticity by dopamine. 4:e09685. Brzosko, Z., Zannone, S., Schultz, W., Clopath, C., and Paulsen, O. (2017). Sequential neuromodu- lation of hebbian plasticity offers mechanism for eLife, page effective reward-based navigation. 6:e27756. Chen, J. X. (2016). The evolution of computing: Alphago. Computing in Science & Engineering, 18(4):4–7. Cho, K. e. a. (2014). Learning phrase repre- sentations using rnn encoder-decoder for sta- arXiv preprint tistical machine translation. arXiv:1406.1078. Compte, A., Brunel, N., Goldman-Rakic, P. S., and Wang, X.-J. (2000). Synaptic mechanisms and network dynamics underlying spatial working memory in a cortical network model. Cerebral Cortex, 10:910–923. Cui, Y., Surpur, C., Ahmad, S., and Hawkins, J. (2015). Continuous online sequence learning with an unsupervised neural network model. CoRR, abs/1512.05463. Frank, M. J., Loughry, B., and O'Reilly, R. C. (2001). Interactions between frontal cortex and basal ganglia in working memory: a computational model. Cognitive, Affective, & Behavioral Neuro- science, 1(2):137–160. 15 Frémaux, N. and Gerstner, W. (2016). Neuromodu- lated spike-timing-dependent plasticity, and the- ory of three-factor learning rules. Frontiers in Neural Circuits, page 9:85. Gers, F. A. and Schmidhuber, J. (2001). Long short- term memory learns context free and context sensitive languages. In Artificial Neural Nets and Genetic Algorithms, pages 134–137. Springer. Gottlieb, J. and Goldberg, M. E. (1999). Activity of neurons in the lateral intraparietal area of the monkey during an antisaccade task. Nature neu- roscience, 2(10):906–912. Graves, A., Wayne, G., I. (2014). Neural turing machines. arXiv preprint arXiv:1410.5401. and Danihelka, Graves, A., Wayne, G., Reynolds, M., Harley, T., Danihelka, I., Grabska-Barwi ´nska, A., Col- menarejo, S. G., Grefenstette, E., Ramalho, T., Agapiou, J., et al. (2016). Hybrid computing using a neural network with dynamic external memory. Nature, 538(7626):471–476. Guerguiev, J., Lillicrap, T. P., and Richards, B. A. (2017). Towards deep learning with segregated dendrites. eLife, page 6:e22901. He, K., Huertas, M., Hong, S. Z., Tie, X., Hell, J. W., Shouval, H., and Kirkwood, A. (2015). Dis- tinct eligibility traces for ltp and ltd in cortical synapses. Neuron, 88(3):528–538. Hochreiter, S. and Schmidhuber, J. (1997). Long computation, Neural short-term memory. 9(8):1735–1780. Legenstein, R., Pecevski, D., and Wolfgang, M. (2008). A learning theory for reward-modulated spike-timing-dependent plasticity with appli- PLOS Comput Biol., cation to biofeedback. 4(10):e1000180. Li, Y. (2017). Deep reinforcement learning: An overview. arXiv preprint arXiv:1701.07274. Mao, T., Kusefoglu, D., Hooks, B. M., Huber, D., Petreanu, L., and Svoboda, K. (2011). Long- range neuronal circuits underlying the interac- tion between sensory and motor cortex. Neuron, 72(1):111–123. Mink, J. W. (1996). The basal ganglia: focused se- lection and inhibition of competing motor pro- grams. Progress in neurobiology, 50(4):381–425. Minsky, M. and Papert, S. (1969). Perceptrons. MIT press. Mnih, V., Kavukcuoglu, K., Silver, D., Rusu, A. A., Veness, J., Bellemare, M. G., Graves, A., Ried- miller, M., Fidjeland, A. K., Ostrovski, G., et al. (2015). Human-level control through deep rein- forcement learning. Nature, 518(7540):529–533. Moore, T. and Armstrong, K. M. (2003). Selective gating of visual signals by microstimulation of frontal cortex. Nature, 421(6921):370–373. Morrison, A., Diesmann, M., and Gerstner, W. (2008). Phenomenological models of synaptic plasticity based on spike timing. Biological Cy- bernetics, 98(6):459–478. Okano, H., Hirano, T., and Balaban, E. (2000). Learning and memory. Proceedings of the National Academy of Sciences, 97(23):12403–12404. O'Reilly, R. C. and Frank, M. J. (2006). Making working memory work: a computational model of learning in the prefrontal cortex and basal gan- glia. Neural computation, 18(2):283–328. Ororbia, A. G., Mikolov, T., and Reitter, D. (2017). Learning simpler language models with the dif- ferential state framework. Neural Computation, 29(12):3327–3352. Pfister, J.-P. and Gerstner, W. (2006). Triplets of spikes in a model of spike timing-dependent Journal of Neuroscience, 26(38):9673– plasticity. 9682. Lillicrap, T. P., Cownden, D., Tweed, D. B., and Ak- erman, C. J. (2016). Random synaptic feedback weights support error backpropagation for deep learning. Nature communications, 7. Roelfsema, P. R. and van Ooyen, A. (2005). Attention-gated reinforcement learning of inter- nal representations for classification. Neural com- putation, 17(10):2176–2214. 16 Roelfsema, P. R., van Ooyen, A., and Watanabe, T. (2010). Perceptual learning rules based on rein- forcers and attention. Trends in cognitive sciences, 14(2):64–71. Waelti, P., Dickinson, A., and Schultz, W. (2001). Dopamine responses comply with basic as- sumptions of formal learning theory. Nature, 412(6842):43–48. Wiering, M. and Schmidhuber, J. (1998). Fast online q(λ). Machine Learning, 33(1):105–115. Xie, X. and Seung, H. S. (2004). Learning in neural networks by reinforcement of irregular spiking. Phys Rev E., 69(4):041909. Yagishita, S., Hayashi-Takagi, A., Ellis-Davies, G. C., Urakubo, H., Ishii, S., and Kasai, H. (2014). A critical time window for dopamine actions on the structural plasticity of dendritic spines. Sci- ence, 345(6204):1616–1620. Rombouts, J. O., Bohte, S. M., and Roelfsema, P. R. (2015). How attention can create synaptic tags for the learning of working memories in sequential tasks. PLOS Computational Biology, 11(3):1–34. Rumelhart, D. E., Hinton, G. E., and Williams, R. J. (1985). Learning internal representations by error propagation. Technical report, California Univ San Diego La Jolla Inst for Cognitive Sci- ence. Samsonovich, A. and McNaughton, B. L. (1997). Path integration and cognitive mapping in a con- tinuous attractor neural network model. Journal of Neuroscience, 17(15):5900–5920. Santoro, A., Bartunov, S., Botvinick, M., Wierstra, D., and Lillicrap, T. (2016). One-shot learn- ing with memory-augmented neural networks. arXiv preprint arXiv:1605.06065. Schultz, W., Apicella, P., and Ljungberg, T. (1993). Responses of monkey dopamine neurons to re- ward and conditioned stimuli during successive steps of learning a delayed response task. Journal of neuroscience, 13(3):900–913. Schultz, W., Dayan, P., and Montague, P. R. (1997). A neural substrate of prediction and reward. Sci- ence, 275(5306):1593–1599. Sutton, R. S. (1984). Temporal Credit Assignment in Reinforcement Learning. PhD thesis. AAI8410337. Sutton, R. S. and Barto, A. G. (1998). Reinforcement learning: An introduction, volume 1. MIT press Cambridge. Tetzlaff, C., Kolodziejski, C., Markelic, I., and Wörgötter, F. (2012). Time scales of memory, learning, and plasticity. Biological Cybernetics, 106(11):715–726. Vasilaki, E., Frémaux, N., Urbanczik, R., Senn, W., and Gerstner, W. (2009). Spike-based reinforce- ment learning in continuous state and action space: When policy gradient methods fail. PLOS Computational Biology, 5(12):e1000586. 17 Supplementary Material Model Validation of AuGMEnT Network: The Saccade-Antisaccade Task The Saccade-AntiSaccade Task (S-AS), presented in the AuGMEnT paper [Rombouts et al. 2015], is inspired by cognitive experiments performed on monkeys to study the memory representations of visual stimuli in the Lateral Intra-Parietal cortex (LIP). The structure of each trial covers different phases in which different cues are presented on a screen and at the end of each episode the agent has to respond accordingly in order to gain a reward. Actually, following a shaping strategy, monkeys received also an intermediate smaller reward when they learnt to fixate on the task-relevant marks at the center of the screen. The details on the procedure of the trials and the experimental results are discussed in Gottlieb and Goldberg [1999]. The agent has to look either to 'Left' (L) or to 'Right' (R) in agreement with a sequence of marks that appear on a screen at each episode. The response, corresponding to the direction of the eye movement (called saccade), depends on the combination of the location and fixation marks. The location cue is a circle displayed either at the left side (L) or at the right side (R) of the screen, while the fixation mark is a square presented at the center that indicates whether the final move has to be concordant with the location cue (Prosaccade - P) or in the opposite direction (Antisaccade - A). As a consequence, there are four types of trials corresponding to the four cue combinations. As can be seen in Figure 1, each trial is struc- tured in five phases: a) start, where the screen is initially empty, b) fix, when the fixation mark appears c) cue, where the location cue is added on the screen, d) delay, in which the location circle disappears for two timesteps, e) go, when the fixation mark vanishes as well and the agent has to give the final response to get the reward. Since the action is given at the end of the trial when the screen is completely empty, the task can be solved only if the network stores and maintains both the Figure 1. Structure of the Saccade-AntiSaccade task. Structure of the trials in all the possible modalities: P-L and A-R have final response L (green arrow), while trials P-R and A-L lead to take action R (red arrow). Figure taken from publication [Rombouts et al. 2015]. stimuli in memory in spite of the delay phase. In addition, the shaping strategy mentioned above is applied in the fix phase of the experiment, by giving an intermediate reward if the agent gazes at the fixation mark for two consecutive timesteps, to ensure that he observes the screen during the whole trial and that the go response is not random but consequential to the cues. So, the reward for the final response is equal to 1.5 units in case of correct response, 0 otherwise, but the intermediate reward for the shaping strategy is smaller, equal to 0.2 units. The most important details about the trial structure are summarized in Table 1. The mean trend of the prediction error during training shows that learning of the S-AS task is achieved by the of AuGMEnT network also in our simulation (Figure 2). In particular, in order to compare it with the reference performance in Rombouts et al. [2015] we applied the same convergence condition, for which training on the S-AS task is considered to be successful if the accuracy for each trial type is higher than 18 Table 1. Table of the S-AS task Task feature Task Structure Inputs Outputs Trial Types Details 5 phases: start, fix, cue, delay, go Fixation mark: Pro-saccade (P) or Anti-saccade (A) Location mark: Left (L) or Right (R) Eye movement: Left (L), Front (F) or Right (R) 1. P+L=L 3. A+L=R 2. P+R=R 4. A+R=L Training dataset Maximum number of trials is 25, 000. Each trial type has equal probability. Correct saccade at go (1.5 units) Fixation of the screen in fix (0.2 units) Rewards Figure 2. Validation of learning perfomance on the saccade task. Decay of the mean predicition error during training of AuGMEnT networks on the S-AS task, computed as the mean number of errors in 50 trials and averaged over 100 simulations. hybrid AuGMEnT and leaky control, proving that they also solve the S-AS task. However, the time to convergence is higher (especially for leaky control) because non-leaky memory is better suited to this simple task. 90% in the last 50 trials. In the original paper, convergence of AuGMEnT is achieved 99.45% of the times, with a training of around 4, 100 trials. In our simulations, the network reaches convergence every time, with a mean time of 2, 063 trials (s.d.= 837.7). The slightly better performance in our simulations could be due to minor differences in the interpretation of the task structure of S-AS or of the convergence criterion, but in any case the error plot in Figure 2 confirms a sharp decrease in the variability of the prediction error after 4, 000 iterations, compliant with the convergence results in the reference paper. In addition, we also show the performance of 19 A Temporal XOR Task B S-AS Task C 12AX Task Figure 3. Spatial representation of temporal XOR, S-AS and 12AX tasks. Schematic representation of the task structures indicating for each cue combination the correct response. Temporal XOR (A) and S-AS (B) tasks have analogous input-output maps and they both have an output space that is not linearly separable. The map of 12AX task (C) is three-dimensional because its structure is based on three hierarchical levels of inputs instead of two. However, for better visualization we add at right the sections of the structure space with respect to the digit inputs that start the outer loops. The output space is still not linearly separable in 3D. 20
1607.06563
1
1607
2016-07-22T06:06:46
Leaky Integrate-and-Fire Neuron under Poisson Stimulation
[ "q-bio.NC" ]
We consider a single Leaky integrate-and-fire neuron stimulated with Poisson process. We develop a method, which allows one to obtain the first passage time probability density function without any additional approximations.
q-bio.NC
q-bio
International Young Scientists Forum on Applied Physics and Engineering YSF-2016 Leaky Integrate-and-Fire Neuron under Poisson Stimulation. Kseniia Kravchuk Synergetics department Bogolyubov institute for Theoretical physics Kyiv, Ukraine [email protected] Abstract -We consider a single Leaky integrate-and-fire neuron stimulated with Poisson process. We develop a method, which allows one to obtain the first passage time probability density function without any additional approximations. Keywords- Leaky Integrate-and-Fire neuron; Poissson process; First Passage Time; rigorous solution I. INTRODUCTION If the input stimulation of a neuron were known, what would be the output? This problem has been addressed many times during the last 5 decades, see e.g. [1]–[3]. A widely used way to deal with this problem is to apply the diffusion approximation. In the diffusion approximation, it is assumed, that each input impulse, received by a neuron, has an infinitesimally small effect on the neuron's state (on its membrane voltage). If so, the number of input impulses needed to trigger the neuron is infinite. Nevertheless, a careful consideration of experimental data shows, that the number of input impulses, needed to trigger a real neuron, is not always sufficiently large to support the diffusion approximation [4]– [6]. Moreover, the diffusion approximation does not allow a consistent theoretical consideration of a neural network, when the output impulses of some neuron are sent to the input of another neuron. That is why, one should cast out the diffusion approximation and consider a more realistic situation, when each input impulse causes a significant change in neuron's membrane voltage, and the number of input impulses, needed to trigger the neuron, is finite. Recently, the first steps were made in this direction for the Binding neuron model [7] and for the Leaky integrate-and-fire (LIF) neuron model [8] stimulated with Poisson stream of input impulses. In [8], the first passage time probability density function (p.d.f.) is derived for the particular case, when 2 impulses are enough to trigger the neuron. Unfortunately, the approach developed in [8] is too complicated and cannot be extended to the general case. In this paper, we develop a general method for the LIF model, which allows to obtain exact analytical results for the first passage time p.d.f. without any additional approximations. II. PROBLEM STATEMENT Consider a separate neuron, stimulated by a sequence of input impulses. Intervals between input impulses are stochastic variables. As input interspike intervals we take Poisson p.d.f. with intensity . the probability distribution for the Assume that any input impulse has a significant effect on neuron's state, and that the amount of impulses needed to trigger the neuron is finite. The neuron is modelled with the LIF model. We find rigorously a probability density P(t) for the first passage time of this system. Namely, we assume that the neuron starts from its resting state at 0 moment of time, and find the probability P(t)dt that the threshold will be reached for the first time at the moment t, with precision dt (t is called the first passage time). A. Leaky Integrate-and-Fire neuronal model At any moment of time, the state of LIF neuron is completely determined by a single real variable V, which resembles the membrane voltage of a real neuron. The resting state is associated with V=0. Each input impulse raises the membrane voltage by a fixed quantity h: V(ti +) = V(ti –)+ h, (1) were ti denotes an arrival time of i-th input impulse, i=1,2,.... Between the membrane voltage decays exponentially with a character time : impulses, input V(ti + t) = V(ti +) e– t/, (2) were t is a time interval free of input impulses. LIF neuron fires an output spike, when the membrane voltage reaches its threshold value V0, and then the neuron comes back to its resting state V=0. III. THE METHOD A. Zones of membrane voltage values Depending on its parameters, LIF neuron may require different number of input impulses to reach the threshold and fire an output spike. Let us denote with N the minimal number of input impulses, needed to trigger the neuron. An expression for N can be easily derived from neuron parameters: N = [V0/h] + 1, (3) where the square brackets denote an integer part of the corresponding expression. In particular realizations of an input stochastic process, the number of input impulses, which trigger the neuron, may be different, depending on impulses' arrival times, but never less than N. Let us divide the range of membrane voltage values, V[0;V0], into zones of height h, beginning from the top, see October 10-14, 2016 Kharkiv, Ukraine 1 International Young Scientists Forum on Applied Physics and Engineering YSF-2016 let us introduce li,i – 1(t) as a flow of probability across the zones' borders due to the exponential decay. Namely, li,i – 1(t) dt gives the probability that at the moment t with precision dt the trajectory of membrane voltage will cross the border of i-th and (i – 1)-th zones due to the exponential decay. In calculation of li,i – 1(t), only trajectories which have never reached the threshold should be accounted. Now, all the transitions between zones can be written as the following system of ODEs: Fig. 1. Zones of membrane voltage values. Fig.1. We obtain (N–1) zones of height h and one more zone of a smaller height at the bottom. The zones can be assigned with numbers, e.g. as at the Fig. 1: 1-st zone: V (0; V0 – (N–1)h), i-th zone: V [V0 – (N–i+1)h; V0 – (N–i)h), i = 2,3,…,N. (4) The resting state, V=0, should be considered as a separate zone, because of nonzero probability, associated with it. So, let us define a 0 zone as follows: 0 zone: V = 0. (5) Let us introduce probabilities p(i,t), i = 0,1,…,N to have the membrane voltage within i-th zone at the moment t, given the neuron didn't reach the threshold before t. We would like to emphasize, that we are not going to find probabilities of the specific values of V at different moments of time – just the probabilities to have the membrane voltage within the certain limits, see (4). Such simplification can be made without any loose of precision in P(t). Indeed, let us imagine, that at some moment of time the membrane voltage takes its value within N-th zone: V[V0 – h; V0). If the neuron obtains an input impulse at this moment, then the membrane voltage will reach its threshold, and an output spike will be generated. The generation of an output impulse will not depend on the exact value of V, since V belongs to the N-th zone. Therefore, output ISI probability density can be found as a product of p(N,t) and the probability to obtain an input impulse within time interval [t;t+dt), which is dt for a Poisson process: P(t)dt = p(N,t)  dt, (6) were  is the intensity of input Poisson process. So, the first passage time problem reduces to finding p(N,t). B. Equations It seems, that the probability p(N,t) should be found together with probabilities p(i,t) for the rest of zones, because the membrane voltage may get across the zone's borders. The mechanisms of such crossings are  jumps,  exponential decay. The first one is activated with the arrival of input impulse, and guarantees, that V will move for one zone up. The second one is activated permanently, and provides the possibility to move for one zone down. In order to account the second mechanism, Here, the terms like p(i,t) in the right-hand side correspond to the jumping out of the i-th zone upward with the arrival of input impulse (see the first mechanism, above). The terms like li,i – 1(t) describe the probability flow across the zones' borders due leak. This system should be supplemented with initial conditions: the exponential to p(0,t)t=0 = 1, p(i,t)t=0 = 0, i = 1,2,…,N. (8) In order to solve (7) and to obtain p(N,t), one should first find expressions for all li,i – 1(t). C. The flow of probability across the zones borders The crossing of a given zones' border at definite time t may happen in different trajectories. Such trajectories may contain a different number k of input impulses, obtained within time interval (0;t), and the arrival times of such impulses may differ as well. That is why, it is natural to find li,i – 1(t) as a sum by k: where functions lk i,i–1(t) correspond to trajectories with k input impulses within interval (0;t). Namely, lk i,i–1(t) dt gives the probability of crossing the border of i-th and (i-1)-th zones at time t with precision dt due to exponential decay (only trajectories which never reached the threshold should be accounted). D. Minimum and maximum number of input impulses The number k of input impulses which contribute to li,i – 1(t) in (9) should be limited both from the top and from the bottom. Indeed, k should be sufficiently large to let the membrane voltage reach the zones' border (which in accordance with (4) is determined by equality V = V0 – (N – i +1) h): (kmin)i = [(V0 – (N – i + 1)h) / h]+ 1 = i – 1, (10) where the definition (3) is taken into account. October 10-14, 2016 Kharkiv, Ukraine 2 International Young Scientists Forum on Applied Physics and Engineering YSF-2016 j is a time interval needed to the membrane voltage to decay from Fig. 2. Ti the value V=jh to the border of i-th and (i-1)-th zones, V=V0–(N–i +1)h. On the other hand, k cannot be arbitrarily large, because otherwise the threshold will be reached before the moment t. This imposes an upper limit for the possible values of k. In order to find such limit, let us first define two sets of time intervals with the following expressions: j is a time interval, needed for the membrane voltage to decay Ti from the level V=jh to the border of i-th and (i-1)-th zones (see Fig. 2), and is the time interval which it takes to the membrane voltage to decay from V0 to the same border. Now, consider the closest packing of input impulses within time t, which does not cause triggering and allows crossing the zones border afterwards, see Fig.3. The first (N–1) impulses may arrive arbitrarily close in time to one another, and they would not be able to trigger the neuron. And the others should be separated with time intervals, needed to the membrane voltage to decay to the level V0–h, to avoid triggering. After the (kmax)i-th impulse there should be a time interval to let the – to the level V = V0 – (N – membrane voltage drop from V = V0 i +1) h of zones' border, see Fig. 3. Therefore, one obtains be rewritten using (11) as follows: , which can Fig. 3. The closest packing of (N+2) input impulses, which does not cause triggering. * – the first (N-1) impulses may arrive arbitrarily close to zero time moment. If t < Ti i-1, the probability flow across the zones' border is zero: li,i–1(t)=0, t < Ti i-1, (14) because any trajectory will not have enough time to reach the i-1, the trajectories may manage to cross the border. If t ≥ Ti zones' border and the probability flow emerges. Therefore, in this case, one should obtain all the lk i,i–1(t) in order to find li,i–1(t). E. Functions lk i,i-1(t) Trajectories contribute to lk i,i–1(t) if the following conditions are satisfied: 1. 2. time interval (0;t) contains k input impulses; trajectory crosses the border between i-th and (i-1)-th zones at time t with precision dt; 3. trajectory haven't reached the threshold until the moment t. Then lk i,i–1(t) will be a measure of trajectories which satisfy conditions (1)–(3), above. This measure can be found as follows. Let us denote with t1, t2,…, tk the arrival times of input impulses in a trajectory which contributes to lk i,i–1(t). The probability to obtain from the Poisson stream k input impulses within time interval (0;t) at the moments t1, t2,…, tk equals ke–t. In order to obtain lk i,i–1(t) one should integrate the this expression over all possible values of arrival times tj: Similarly, if , (kmax)i can be obtained as (kmax)i = i –1 + j, t  [Ti i-1+j; Ti i+j )j, j = 0,1,…, N – i –1 (13) Finally, we have obtained, that (kmax)i increases with the increase in t and changes its value by unity at the points and the corresponding values of (kmax)i are i–1, i,…, N–1, N, N+1, N+2, N+3, … where the integration domain T is defined by conditions 2 and 3, above. In (15), the integration over t1 is missing, because the variables t1, t2,…, tk are not independent. Indeed, condition 2 ensures that one of the arrival times can be found using all the others, e.g. one can find t1 as a function of t2,…, tk and t. Domain T in the space with coordinates t2,…, tk has a complicated structure, due to the conditions 2 and 3, above. This problem may be simplified by appropriate variable change. Let us introduce variables V2 , V3 ,…, Vk, where Vj , j=2,…,k, denotes the membrane voltage just after the arrival of j-th input impulse: October 10-14, 2016 Kharkiv, Ukraine 3 International Young Scientists Forum on Applied Physics and Engineering YSF-2016 Vj  V(tj +), j = 2, 3, …, k. (16) were is defined by the following expressions: The mapping t2,…,tk  V2,…,Vk is one-to-one, so we can perform the corresponding variable change in (15): where the integration domain V in a space with coordinates V2,…, Vk corresponds to the domain T in a space with coordinates t2,…, tk and is defined by the conditions 2 and 3, above. The Jacobean in (17), can be easily found using (1), (2) and (16): Such variable change makes it possible to find upper limits of integration easily. Indeed, the first N–1 impulses may be located arbitrarily close to each other, therefore, Vj < jh for j = 2,3,…, N – 1. Then, after the N-th input impulse, the threshold can be reached and the condition 3 should be accounted. In terms of the membrane voltage, this condition can be written as the simple inequality Vj < V0, j = N, N+1,…, k. Finally, one can write formally: were the upper limits are and the lower limits Vj are still to be found using the conditions 2 and 3, above. In order to find Vj, one should rewrite the condition 2 in terms of membrane voltage variables Vj, using (1) and (2): The lower limit Vj is achieved, if all the previous impulses, m = 1,…, j–1, are located as early as possible, given the conditions , m = 1,2,…,j–1, the 1–3, above. Let us denote with membrane voltage after the arrival of m-th input impulse in such arrangement. The first (N–1) impulses may be located as close as possible to the moment 0, tm = 0, m=1,2,…, N–1, (22) and the others should be separated with time intervals, needed to membrane voltage to decay to the level V0–h, as at Fig. 3. That is why, So, finally, in order to obtain the first passage time p.d.f., one should 1) find all the lk i,i–1(t), using expressions (17)–(20), (24) and (25); 2) for any fixed i, sum up all the lk i,i–1(t) and obtain the probability flow li,i–1(t) across the zones' border, using eqs. (9), (10), (12) and (13); 3) solve the system (7) of ODEs for the probabilities p(i,t) to have the membrane voltage within i-th zone; 4) having p(N,t), obtain expression for the first passage time p.d.f., using (6). IV. CONCLUSSIONS In a present paper, we have considered a single leaky integrate-and-fire neuron stimulated with Poisson stream of input impulses. We have developed a method, which allows one to obtain rigorous expressions for the first passage time p.d.f. for such neuron. The expressions obtained will be different for different relations between neuron's threshold and the altitude of input impulse in the LIF model. In a future, we intend to apply the developed method to a particular case of such relation and to obtain specific expressions for the first passage time p.d.f. in this case. REFERENCES [1] P.I.M. Johannesma, "Diffusion models for the stochastic activity of neurons" in Neural networks, R. Caianiello, Ed. Berlin: Springer, pp. 116-144, 1968. [2] L.M. Ricciardi "Diffusion Processes and Related Topics in Biology" in Lecture Notes in Biomathematics, Vol. 14, Berlin: Springer, 1977. [3] S. Ditlevsen, P. Lansky, "Parameters of stochastic diffusion processes estimated from observations of first-hitting times: Application to the leaky integrate-and-fire neuronal model", Phys. Rev. E., Vol. 76, p. 041906, 2007. [4] R. Miles, "Synaptic excitation of inhibitory cells by single CA3 hippocampal pyramidal cells of the guinea-pig in vitro", J. Physiol., vol. 428, pp. 61-77, September 1990. [5] B. Barbour, "Synaptic currents evoked by purkinje cells by stimulating individual granule cells", Neuron, vol. 11, no. 4, pp. 759-769, October 1993. [6] P. Andersen, "Synaptic integration in hyppocampal neurons" in Fidia Research Foundation Neuroscience Award Lectures, New York: Raven Press Ltd., 1991, pp. 51-71. Substituting values of from (23) and t1=0 from (22) to (21), one obtains the lower limit Vj as a function of Vj+1,…,Vk: October 10-14, 2016 Kharkiv, Ukraine [7] A.K. Vidybida, "Output stream of a binding neuron", Ukrainian Mathematical Journal, Springer US, vol. 59, no. 12, pp. 1819-1839, 2008. [8] A.K. Vidybida, "Output stream of leaky integrate-and-fire neuron", Reports of the NAS of Ukraine, vol. 12, pp. 18-23, 2014. 4
1209.3411
1
1209
2012-09-15T14:28:54
A Computational Model of the Effects of Drug Addiction on Neural Population Dynamics
[ "q-bio.NC", "cs.SI" ]
Reward processing and derangements thereof, such as drug addiction, involve the coordinated activity of many brain areas. Prior work has identified many behavioral, molecular biological and single neuron changes throughout the mesocorticolimbic system that reflect and drive addictive behavior. Subpopulations in the ventral tegemental area (VTA) encode positive reward prediction error, negative reward prediction error, and the magnitude of the reward. Phasic activity in VTA dopaminergic neurons correlates with hedonic value. Tonic activity of groups in the dorsomedial prefrontal cortex (dmPFC) can encode antidepressant states. However, little is known about how drug addiction might affect population encoding across larger brain regions. Here, we compare the information content associated with network patterns in naive, acutely intoxicated and chronically addicted states in a plastic attractor network. We found that addiction decreases the network's ability to store and discriminate among patterns of activity. Altered dopaminergic tone flattens the energy landscape and decreases the entropy associated with each network pattern. Altered dmPFC activity produces signal-to-noise deficits similar to computational models of schizophrenia. Our results provide a conceptual framework for interpreting altered neural population dynamics in psychopathological states based on information theory. They also suggest a view of the subtypes of depression as on a continuum of combinations of cortical and subcortical dysfunction. This suggests that patients who suffer from depression with psychotic features will have more cortical than mesolimbic dysfunction. Furthermore, our framework can be applied to other psychiatric illnesses and so may help us, in general, quantitatively understand psychiatric illnesses as disorders in the representation and processing of information by distributed brain networks.
q-bio.NC
q-bio
Model Evolution of Voltage over Time Consider a group of interacting populations of neurons. Equation 1 describes the evolution over time of the membrane voltages for the ith ensemble. Table 1 explains the remaining variables. τi d~vi dt = −~vi + Mii~vi + X i6=j Mij~vj + Wi~u Symbol Meaning Mij Wi ~u τ Connections from jth ensemble to ith Weight of input for ith ensemble Input Time constant Table 1: Meaning of symbols in Equation 1 To represent n ensembles, we may combine n versions of Equation 1, as Equation 2 describes. τv d~v dt = −~v + M~v + W~u (1) (2) ~v =   ~vi ~vj ... ~vn   , M =   Mii Mij Mji Mjj ... Mni . . . Min . . . Mnn   If the real part of the eigenvalues of M are less than one, then the system will evolve to the voltage that Equation 3 describes. ~v∞ = KW~u K = (I − M)−1 (3) Islands of Steady State Behavior M need not be symmetric. But, if some block matrices within it are, the corresponding subpopulations can approach their own steady state even if the network is still unstable. If, furthermore, the combination of recurrent and feedforward input to those subpopulations is a saturating function, then the system's Lyapunov function is bounded and fixed points for that subpopulation must exist (Cohen and Grossberg, 1983). One reasonable way to create a symmetric weight matrix for Mk, the block matrix that describes the kth ensemble is to assume that it recognizes one of N memory patterns, (cid:8)v1, v2, . . . , vm, . . . , vN(cid:9). Assume that a subpopulation with n neurons signals its recognition of any memory pattern, ~vm, by displaying a voltage vector c~vm. One matrix that accomplishes this is: M = λ c2αN (1 − α) N X (~v − αcn) ⊗ (~v − αcn) − n ⊗ n αN (4) Longer-term Effects on Neural Population Dynamics Assume that two additional processes occur as defined by Equations 5 and 6. Both are much slower than Equation 2. One modifies W and an even slower one modifies M. τW dW dt = h~v~ui − α h~v~vi W α > 0 τM dM dt = − (W~u) ~v + K−1 (5) (6) Replacing ~v with ~v∞in Equations 5 and 6 yields Equations 7 and 8. Note the appearance of the autocorrelation matrices for the stimulus, Q, and network activity, R. τW dW dt = KWQ − αR∞W Q = h~u~ui , R∞ = h ~v∞ ~v∞i (7) 1 τM dM dt = − (W~u) (KW~u) + K−1 (8) Equations 7 and 8 describe somewhat contrasting behaviors. Equation 7 aligns the correlation structure of the network activity with the strength of its recurrent connections. Equation 8 urges the outputs to be decorrelated. A faster correlating influence and slower decorrelating one allow oscillations in the correlation of network activity. According to this model, interestingly, those oscillations are dependent on input but not directly on the correlation structure of the input. Remarks on the Structure of the Model There is an interesting concordance between R and Equation 4. If Rmem = h~vmem~vmemi, then M ∝ P Rmem. This further highlights how interrelated Equations 7 and 8 are. Effect of Correlation Structure on the Dynamics of the Feedforward Weights If the stimulus is random, that is Q = I, then the feedforward weights, W, stop changing only when K = αR∞. Com- bining the definition of K in Equation 3 with the observation that R∞ must haverank 1 we note that such stimuli prevent this system from recognizing any pattern that requires a distribution of activities over the network. Moreover, consider- ing the rank-nullity theorem and ranks of Q and R, one can see that this result holds for any stimulus autocorrelation matrix, Q, that results from the outer product of a vector with itself. It is next natural to consider how this system responds to many superimposed stimuli that each have different corre- lation structures. That is, consider aQ that results from the sum of i correlation matrices. Each of those autocorrelation matrices results from the outer product of the ith activity pattern with itself, as Equation 9 describes. Q = X i h~ui~uii (9) If we assume that all the input patterns are pairwise independent, then the rank of Q becomes equal to i, whose upper bound we assume to be the number of neurons. This stands in contrast to the result of the previous paragraph. Effects of Correlation Structure on the Dynamics of Network Connections By studying Equation 8 we can gain some insight into how ~u and W influence M. If ~u lies in the nullspace of the columns of W, then M approaches the identity matrix. Said another way, when ~u and W jam each other,M tries to span the largest basis it can. The contrary effects of Nonrandom Constant Stimulus Consider any brief nonrandom stimulus such that, Q 6= I but is constant. ThenW only stops changing if the columns of Q are in the nullspace of the rows of W. BecauseQ results from the tensor product of a vector with itself, it has rank 1 and so it may lie in the nullspace of W. Brief Pulse If a stimulus is presented to the system for a brief period of time, it will only cause sustained activity in the system if the activity that it induces is itself a fixed point of the system. That is, for a stimulus, ~u, and its response, ~v, Equation must hold. Extension to Drug Addiction Craving . Similar response to variable reinforcement and huge rewards. Let the ith subpopulation of Equation 2, in analogy with a proposed role for the dopaminergic neurons in ventral tegmental area (VTA) represent the positive expecrted reward of a stimulus. Let another population, the jth one, in analogy with the GABAergic neurons in the VTA represent the negative expected reward. 2
1610.00161
3
1610
2017-04-07T18:45:30
Towards deep learning with segregated dendrites
[ "q-bio.NC" ]
Deep learning has led to significant advances in artificial intelligence, in part, by adopting strategies motivated by neurophysiology. However, it is unclear whether deep learning could occur in the real brain. Here, we show that a deep learning algorithm that utilizes multi-compartment neurons might help us to understand how the brain optimizes cost functions. Like neocortical pyramidal neurons, neurons in our model receive sensory information and higher-order feedback in electrotonically segregated compartments. Thanks to this segregation, the neurons in different layers of the network can coordinate synaptic weight updates. As a result, the network can learn to categorize images better than a single layer network. Furthermore, we show that our algorithm takes advantage of multilayer architectures to identify useful representations---the hallmark of deep learning. This work demonstrates that deep learning can be achieved using segregated dendritic compartments, which may help to explain the dendritic morphology of neocortical pyramidal neurons.
q-bio.NC
q-bio
Towards deep learning with segregated dendrites Jordan Guergiuev1,2, Timothy P. Lillicrap4, and Blake A. Richards1,2,3,* 1 Department of Biological Sciences, University of Toronto Scarborough, Toronto, ON, Canada 2 Department of Cell and Systems Biology, University of Toronto, Toronto, ON, Canada 3 Learning in Machines and Brains Program, Canadian Institute for Advanced Research, Toronto, ON, Canada 4 DeepMind Technologies Inc., London, UK * Corresponding author, email: [email protected] Abstract Deep learning has led to significant advances in artificial intelligence, in part, by adopting strategies motivated by neurophysiology. However, it is unclear whether deep learning could occur in the real brain. Here, we show that a deep learning algorithm that utilizes multi-compartment neurons might help us to understand how the brain optimizes cost functions. Like neocortical pyramidal neurons, neurons in our model receive sensory information and higher-order feedback in electrotonically segregated compartments. Thanks to this segregation, the neurons in different layers of the network can coordinate synaptic weight updates. As a result, the network can learn to categorize images better than a single layer network. Furthermore, we show that our algorithm takes advantage of multilayer architectures to identify useful representations-the hallmark of deep learning. This work demonstrates that deep learning can be achieved using segregated dendritic compartments, which may help to explain the dendritic morphology of neocortical pyramidal neurons. Introduction Deep learning refers to an approach in artificial intelligence (AI) that utilizes neural networks with multiple layers of processing units. Importantly, deep learning algorithms are designed to take advantage of these multi-layer network architectures in order to generate hierarchical representations wherein each successive layer identifies increasingly abstract, relevant variables for a given task (Bengio and LeCun, 2007; LeCun et al., 2015). In recent years, deep learning has revolutionized machine learning, opening the door to AI applications that can rival human capabilities in pattern recognition and control (Mnih et al., 2015; Silver et al., 2016; He et al., 2015). Interestingly, the representations that deep learning generates resemble those observed in the neocortex, particularly in higher-order sensory areas (Kubilius et al., 2016; Khaligh-Razavi and Kriegeskorte, 2014; Cadieu et al., 2014), suggesting that something akin to deep learning is occurring in the mammalian brain (Yamins and DiCarlo, 2016; Marblestone et al., 2016). Yet, a large gap exists between deep learning in AI and our current understanding of learning and memory in neuroscience. In particular, unlike deep learning researchers, neuroscientists do not yet have a solution to the "credit assignment problem" (Rumelhart et al., 1986; Lillicrap et al., 2016; Bengio et al., 2015). The credit assignment problem refers to the fact that the behavioral consequences of synaptic changes in early layers of a neural network depend on the current connections in the downstream layers. For example, consider the behavioral effects of synaptic changes, i.e. long-term potentiation/depression 1/41 (LTP/LTD), occurring between different sensory circuits of the brain. Exactly how these synaptic changes will impact behavior and cognition depends on the downstream connections between the sensory circuits and motor or associative circuits (Figure 1A). Hence, learning to optimize some behavioral or cognitive function requires a method for identifying what a given neuron's place is within the wider neural network, i.e. it requires a means of assigning "credit" (or "blame") to neurons in early sensory areas for their contribution to the final behavioral output (LeCun et al., 2015; Bengio et al., 2015). Despite its importance for real-world learning, the credit assignment problem has received little attention in neuroscience. The lack of attention to credit assignment in neuroscience is, arguably, a function of the history of biological studies of synaptic plasticity. Due to the well-established dependence of LTP and LTD on presynaptic and postsynaptic activity, current theories of learning in neuroscience tend to emphasize Hebbian learning algorithms (Dan and Poo, 2004; Martin et al., 2000), that is, learning algorithms where synaptic changes depend solely on presynaptic and postsynaptic activity. Hebbian learning models can produce representations that resemble the representations in the real brain (Zylberberg et al., 2011; Leibo et al., 2017) and they are backed up by decades of experimental findings (Malenka and Bear, 2004; Dan and Poo, 2004; Martin et al., 2000). But, current Hebbian learning algorithms do not solve the credit assignment problem, nor do global neuromodulatory signals used in reinforcement learning (Lillicrap et al., 2016). As a result, deep learning systems from AI that can do credit assignment outperform existing Hebbian models of perceptual learning on a variety of tasks (Yamins and DiCarlo, 2016; Khaligh-Razavi and Kriegeskorte, 2014). This suggests that a critical, missing component in our current models of the neurobiology of learning and memory is an explanation of how the brain solves the credit assignment problem. However, the most common solution to the credit assignment problem in AI is to use the back- propagation of error algorithm (Rumelhart et al., 1986). Backpropagation assigns credit by explicitly using current downstream synaptic connections to calculate synaptic weight updates in earlier layers, commonly termed "hidden layers" (LeCun et al., 2015) (Figure 1B). This technique, which is sometimes referred to as "weight transport", involves non-local transmission of synaptic weight information between layers of the network (Lillicrap et al., 2016; Grossberg, 1987). Weight transport is clearly unrealistic from a biological perspective (Bengio et al., 2015; Crick, 1989). It would require early sensory processing areas (e.g. V1, V2, V4) to have precise information about billions of synaptic connections in downstream circuits (MT, IT, M2, EC, etc.). According to our current understanding, there is no physiological mechanism that could communicate this information in the brain. Some deep learning algorithms utilize purely Hebbian rules (Scellier and Bengio, 2016; Hinton et al., 2006). But they depend on feedback synapses that are symmetric with feedforward synapses to solve the credit assignment problem (Scellier and Bengio, 2016; Hinton et al., 2006), which is essentially a version of weight transport. Altogether, these artificial aspects of current deep learning solutions to credit assignment have rendered many scientists skeptical to the proposal that deep learning occurs in the real brain (Crick, 1989; Grossberg, 1987; Harris, 2008; Urbanczik and Senn, 2009). Recent findings have shown that these problems may be surmountable, though. Lillicrap et al. (2016), Lee et al. (2015) and Liao et al. (2015) have demonstrated that it is possible to solve the credit assignment problem even while avoiding weight transport or symmetric feedback weights. The key to these learning algorithms is the use of feedback signals from output layers that are transmitted to calculate a local error signal, which then guides synaptic updates in hidden layers (Lee et al., 2015; Lillicrap et al., 2016; Liao et al., 2015). These learning algorithms can take advantage of multi-layer architectures, leading to performance that rivals backpropagation (Lee et al., 2015; Lillicrap et al., 2016; Liao et al., 2015). Hence, this work has provided a significant breakthrough in our understanding of how the real brain might do credit assignment. Nonetheless, the models of Lillicrap et al. (2016), Lee et al. (2015) and Liao et al. (2015) involve some problematic assumptions. Specifically, although it is not directly stated in all of the papers, there is an implicit assumption that there is a separate feedback pathway for transmitting the signals that drive synaptic updates in hidden layers (Figure 2A). Such a pathway is required in these models because the synaptic weight updates in the hidden layers depend on the difference between feedback that is 2/41 generated in response to a purely feedforward propagation of sensory information, and feedback that is guided by a teaching signal (Lillicrap et al., 2016; Lee et al., 2015; Liao et al., 2015). In order to calculate this difference, sensory information must be transmitted separately from the feedback signals that are used to drive learning. In single compartment neurons, keeping feedforward sensory information separate from feedback signals is impossible without a separate pathway. At face value, such a pathway is possible. But, closer inspection uncovers a couple of difficulties with such a proposal. First, the error signals that solve the credit assignment problem are not global error signals (like neuromodulatory signals used in reinforcement learning). Rather, they are cell-by-cell error signals. This would mean that the feedback pathway would require some degree of pairing, wherein each neuron in the hidden layer is paired with a feedback neuron (or circuit). That is not impossible, but there is no evidence to date of such an architecture in the neocortex. Second, the error that is communicated to the hidden layer is signed (i.e. it can be positive or negative), and the sign determines whether LTP or LTD occur in the hidden layer neurons (Lee et al., 2015; Lillicrap et al., 2016; Liao et al., 2015). Communicating signed signals with a spiking neuron can theoretically be done by using a baseline firing rate that the neuron can go above (for positive signals) or below (for negative signals). But, in practice, such systems are difficult to operate because as the error gets closer to zero any noise in the spiking of the neuron can switch the sign of the signal, which switches LTP to LTD, or vice versa. This means that as learning progresses the network's ability to communicate error signs gets worse. Therefore, the real brain's specific solution to the credit assignment problem is unlikely to involve a separate feedback pathway for cell-by-cell, signed signals to instruct plasticity. However, segregating the integration of feedforward and feedback signals does not require a separate pathway if neurons have more complicated morphologies than the point neurons typically used in artificial neural networks. Taking inspiration from biology, we note that real neurons are much more complex than single-compartments, and different signals can be integrated at distinct dendritic locations. Indeed, in the primary sensory areas of the neocortex, feedback from higher-order areas arrives in the distal apical dendrites of pyramidal neurons (Manita et al., 2015; Budd, 1998; Spratling, 2002), which are electrotonically very distant from the basal dendrites where feedforward sensory information is received (Larkum et al., 1999, 2007, 2009). Thus, as has been noted by previous authors (Kording and Konig, 2001; Spratling, 2002; Spratling and Johnson, 2006), the anatomy of pyramidal neurons may actually provide the necessary segregation of feedforward and feedback information to calculate local error signals and perform deep learning in biological neural networks. Here, we show how deep learning can be implemented if neurons in hidden layers contain segregated "basal" and "apical" dendritic compartments for integrating feedforward and feedback signals separately (Figure 2B). Our model builds on previous neural networks research (Lee et al., 2015; Lillicrap et al., 2016) as well as computational studies of supervised learning in multi-compartment neurons (Urbanczik and Senn, 2014; Kording and Konig, 2001; Spratling and Johnson, 2006). Importantly, we use the distinct basal and apical compartments in our neurons to integrate feedback signals separately from feedforward signals. With this, we build a local target for each hidden layer that coordinates learning across the layers of the network. We demonstrate that even with random synaptic weights for feedback into the apical compartment, our algorithm can coordinate learning to achieve classification of the MNIST database of hand-written digits better than can be achieved with single layer networks. Furthermore, we show that our algorithm allows the network to take advantage of multi-layer structures to build hierarchical, abstract representations, one of the hallmarks of deep learning (LeCun et al., 2015). Our results demonstrate that deep learning can be implemented in a biologically feasible manner if feedforward and feedback signals are received at electrotonically segregated dendrites, as is the case in the mammalian neocortex. Results A network architecture with segregated dendritic compartments Deep supervised learning with local weight updates requires that each neuron receive signals that can be used to determine its "credit" for the final behavioral output. We explored the idea that the 3/41 cortico-cortical feedback signals to pyramidal cells could provide the required information for credit assignment. In particular, we were inspired by four observations from both machine learning and biology: 1. Current solutions to credit assignment without weight transport require segregated feedforward and feedback signals (Lee et al., 2015; Lillicrap et al., 2016). 2. In the neocortex, feedforward sensory information and higher-order cortico-cortical feedback are largely received by distinct dendritic compartments, namely the basal dendrites and distal apical dendrites, respectively (Spratling, 2002; Budd, 1998). 3. The distal apical dendrites of pyramidal neurons are electrotonically distant from the soma, such that passive transmission from the distal dendrites to the soma is significantly attenuated. Instead, apical communication to the soma depends on active propagation through the apical dendritic shaft driven by voltage-gated calcium channels. These non-linear, active events in the apical shaft generate prolonged upswings in the membrane potential, known as "plateau potentials", which can drive burst firing at the soma (Larkum et al., 1999, 2009). 4. Plateau potentials driven by apical activity can guide plasticity in vivo (Bittner et al., 2015). With these considerations in mind, we hypothesized that the computations required for credit assignment could be achieved without any separate pathways for feedback signals. Instead, they could be achieved by having two distinct dendritic compartments in each hidden layer neuron: a "basal" compartment, strongly coupled to the soma for integrating feedforward information, and an "apical" compartment for integrating feedback information that would only drive activity at the soma when "plateau potentials" occur (Figure 3A). As an initial test of this concept we built a network with a single hidden layer. Although this network is not very "deep", even a single hidden layer can improve performance over a one-layer architecture if the learning algorithm solves the credit assignment problem (Bengio and LeCun, 2007; Lillicrap et al., 2016). Hence, we wanted to initially determine whether our network could take advantage of a hidden layer to reduce error at the output layer. The network architecture is illustrated in Figure 3A. An image from the MNIST data set is used to set the spike rates of (cid:96) = 784 Poisson point-process neurons in the input layer (one neuron per image pixel, rates-of-fire determined by pixel intensity). These project to a hidden layer with m = 500 neurons. The neurons in the hidden layer are composed of three distinct compartments with their own voltages: the apical compartments (with voltages described by the vector A(t) = [A1(t), ..., Am(t)]), the basal compartments (with voltages B(t) = [B1(t), ..., Bm(t)]), and the somatic compartments (with voltages C(t) = [C1(t), ..., Cm(t)]). (Note: for notational clarity, all vectors and matrices in the paper are in boldface.) The voltages in the dendritic compartments are calculated as weighted sums of a postsynaptic potential kernel convolved with the incoming spike train (see Materials and Methods, equations (10) and (11)), while the somatic voltages are calculated as leaky integrators of the dendritic inputs, i.e. for the ith hidden layer neuron: dCi(t) dt = −gLCi(t) + gB(Bi(t) − Ci(t)) + gA(Ai(t) − Ci(t)) (1) where gL, gB and gA represent the leak conductance, the conductance from the basal dendrites, and the conductance from the apical dendrites, respectively. Note, for mathematical simplicity we are assuming a resting membrane potential of 0 V and a membrane capacitance of 1 F (these values do not affect the results). We implement electrotonic segregation in the model by altering the gA value-low values for gA lead to electrotonically segregated apical dendrites. In the initial set of simulations we set gA = 0, but we relax this hard constraint in later simulations. The somatic compartments generate spikes using Poisson processes. The instantaneous rates of these processes are described by the vector 4/41 λC(t) = [λC m(t)], which is in units of spikes/s or Hz. These rates-of-fire are determined by a non-linear sigmoid function, σ(·), applied to the somatic voltages, i.e. for the ith hidden layer neuron: 1 (t), ..., λC λC i (t) = λmaxσ(Ci(t)) = λmax 1 1 + e−Ci(t) (2) where λmax is the maximum rate-of-fire for the neurons. Spiking inputs from the input layer arrive at the basal compartments via the m × (cid:96) synaptic weight matrix W 0, hidden layer somatic spikes are projected to the output layer neurons via the n× m synaptic weight matrix W 1, and spiking inputs from the output layer arrive at the apical compartments via the m × n synaptic weight matrix Y (Figure 3A). The output layer neurons consist of n = 10 two compartment neurons (one for each image category), similar to those used in a previous model of dendritic prediction learning (Urbanczik and Senn, 2014). The output dendritic voltages (V (t) = [V1(t), ..., Vn(t)]) and somatic voltages (U (t) = [U1(t), ..., Un(t)]) are updated in a similar manner to the hidden layer basal compartment and soma, respectively (see Materials and Methods, equations (14) and (15)). As well, like the hidden layer neurons, the output neurons spike using Poisson processes whose instantaneous rates, λU (t) = [λU n (t)], are determined by the somatic voltages, i.e. λU i (t) = λmaxσ(Ui(t)). Unlike the hidden layer, however, the output layer neurons also receive a set of "teaching signals", specifically, excitatory and inhibitory conductances that push their voltages towards target values (see Materials and Methods, equations (15) and (16)). These teaching signals are similar to the targets that are used for training in deep artificial neural networks (Bengio and LeCun, 2007; LeCun et al., 2015). Whether any such teaching signals exist in the real brain is unknown, though there is evidence that animals can represent desired behavioral outputs with internal goal representations (Gadagkar et al., 2016). 1 (t), ..., λU Critically, we define two different modes of integration in the hidden layer neurons: "transmit" and "plateau". During the transmit mode, the apical compartment is considered to be electrotonically segregated from the soma so it affects the somatic voltage minimally (depending on gA), and most of the conductance to the soma comes from the basal compartment (Figure 3B, left). In contrast, during the plateau mode, the apical voltage is averaged over the most recent 20-30 ms period and the sigmoid non-linearity is applied to it, giving us "plateau potentials" (see equation (3) below). These non-linear versions of the apical voltages are then transmitted to the somatic and basal compartments for synaptic updates (Figure 3B, right). The intention behind this design was to mimic the non-linear transmission from the apical dendrites to the soma that occurs during a plateau potential driven by calcium spikes in the apical dendritic shaft (Larkum et al., 1999). Furthermore, the temporal averaging was intended to mimic, in part, the temporal dynamics introduced by NMDA plateaus (Schiller et al., 2000), which are particularly good for driving apical calcium spikes (Larkum et al., 2009). To train the network we alternate between two phases. First, we present an image to the input layer without any signals to the output layer during the "forward" phase, which occurs between times t0 to t1. At t1 a plateau potential occurs in all the hidden layer neurons and the "target" phase begins. During this phase the image continues to drive the input layer, but now the output layer also receives teaching conductances, which force the output units closer to the correct answer. For example, if an image of a '9' is presented, then over the time period t1-t2 the '9' neuron in the output layer receives strong excitatory conductances from the teaching signal, while the other neurons receive strong inhibitory conductances (Figure 3C). This phase lasts from t1 to t2, and at t2 another set of plateau potentials occurs in the hidden layer neurons. The result is that we have plateaus potentials in the hidden layer neurons for both the forward (αf = [αf m]) and target (αt = [αt 1 , ..., αf 1, ..., αt m]) phases: 5/41 (cid:90) t1 (cid:90) t2 t1−tp t2−tp αf i = σ( αt i = σ( Ai(t)dt) Ai(t)dt) (3) where tp is the integration time for the plateau potential (see Materials and Methods). Note that since the plateau potentials are generated using the same sigmoid non-linearity that we use to calculate the rate-of-fire at the soma, they are analogous to the firing rates of the hidden layer neurons. This allows us to use the plateau potentials to define target firing rates for the neurons (see below). The network is simulated in near continuous-time (except that each plateau is considered to be instantaneous), and the intervals between plateaus are randomly sampled from an inverse Gaussian distribution (Figure 3D, top). As such, the specific amount of time that the network is presented with each image and teaching signal is stochastic, though usually somewhere between 50-60 ms of simulated time (Figure 3D, bottom). This stochasticity was not necessary, but it demonstrates that although the system operates in phases, the specific length of the phases is not important as long as they are sufficiently long to permit integration (see Lemma 1). In the data presented in this paper, all 60,000 images in the MNIST training set were presented to the network one at a time, and each exposure to the full set of images was considered an "epoch" of training. At the end of each epoch, the network's classification error rate on a separate set of 10,000 test images was assessed with a single forward phase (see Materials and Methods). It is important to note that there are many aspects of this design that are not physiologically accurate. Most notably, stochastic generation of synchronized plateau potentials across a population is not an accurate reflection of how real pyramidal neurons operate, since apical calcium spikes are determined by a number of concrete physiological factors in individual cells, including back-propagating action potentials, spike-timing and inhibitory inputs (Larkum et al., 1999, 2007, 2009). However, we note that calcium spikes in the apical dendrites can be prevented from occurring via the activity of distal dendrite targeting inhibitory interneurons (Murayama et al., 2009), which can synchronize pyramidal activity (Hilscher et al., 2017). Furthermore, distal dendrite targeting neurons can themselves can be rapidly inhibited in response to temporally precise neuromodulatory inputs (Pi et al., 2013; Pfeffer et al., 2013; Karnani et al., 2016; Hangya et al., 2015; Brombas et al., 2014). Therefore, it is entirely plausible that neocortical micro-circuits would generate synchronized pyramidal plateaus at punctuated periods of time in response to disinhibition of the apical dendrites governed by neuromodulatory signals that determine "phases" of processing. Alternatively, oscillations in population activity could provide a mechanism for promoting alternating phases of processing and synaptic plasticity (Buzs´aki and Draguhn, 2004). But, complete synchrony of plateaus in our hidden layer neurons is not actually critical to our algorithm-only the temporal relationship between the plateaus and the teaching signal is critical (see below). This relationship itself is arguably plausible given the role of neuromodulatory inputs in dis-inhibiting the distal dendrites of pyramidal neurons (Karnani et al., 2016; Brombas et al., 2014). Of course, we are engaged in a great deal of speculation here. But, the point is that our model utilizes anatomical and functional motifs that are analogous to what is observed in the neocortex. Importantly for the present study, the key issue is the use of segregated dendrites and distinct transmit and plateau modes to solve the credit assignment problem. Credit assignment with segregated dendrites To solve the credit assignment problem without using weight transport, we had to define a local error function for the hidden layer that somehow takes into account the impact that each hidden layer neuron has on the output at the final layer. In other words, credit assignment could only be achieved if we knew that changing a hidden layer synapse to reduce the local error would also help to reduce error at the output layer. To obtain this guarantee, we defined local targets for the output and the hidden layer, i.e. desired firing rates for both the output layer neurons and the hidden layer neurons. Learning is then 6/41 a process of changing the synaptic connections to achieve these target firing rates across the network. Importantly, we defined our hidden layer targets in a similar manner to Lee et al. (2015). These hidden layer targets coordinate with the output layer by incorporating the feedback information contained in the plateau potentials. In this way, credit assignment is accomplished through mechanisms that are spatially local to the hidden layer neurons. Specifically, to create the output and hidden layer targets, we use the average rates-of-fire and average voltages during different phases. We define: y X = [X1 y , ..., Xk y ] (cid:90) t1 (cid:90) t2 t1−tp t2−tp f Xi = t Xi = Xi(t)dt Xi(t)dt (4) where y ∈ {f, t}, k ∈ {(cid:96), m, n} and Xi(t) can be a voltage variable (e.g. Ci(t) or Ui(t)) or a rate-of- i (t)). Using the averages over different phases, we then define the target n ], to be the average rates-of-fire during the target fire variable (e.g. λC rates-of-fire for the output layer, λU = [λU phase: i (t) or λU 1 , ..., λU t λU i = λU i (5) For the hidden layer we define the target rates-of-fire, λC = [λC m], using the average rates-of-fire during the forward phase and the difference between the plateau potentials from the forward and transmit phase: 1 , ..., λC f λC i = λC i + αt i − αf i (6) The goal of learning in the hidden layer is to change the synapses W 0 to achieve these targets in response to the given inputs. More generally, for both the output layer and the hidden layer we then define error functions, L1 and L0, respectively, based on the difference between the local targets and the activity of the neurons during the forward phase (Figure 4A): L1 = λU − λmaxσ(U L0 = λC − λmaxσ(C f f 2 )2 )2 2 (7) Note that the loss function for the hidden layer, L0, will, on average, reduce to the difference αt − αf2 2. This provides an intuitive reason for why these targets help with credit assignment. Specifically, a hidden layer neuron's error is small only when the output layer is providing similar feedback to it during the forward and target phases. Put another way, hidden layer neurons "know" that they are "doing well" when they receive the same feedback from the output layer regardless of whether or not the teaching signal is present. Thus, hidden layer neurons have access to some information about how they are doing in helping the output layer to achieve its targets. More formally, it can be shown that the output and hidden layer error functions will, on average, agree with each other, i.e. if the hidden layer error is reduced then the output layer error is also reduced. Specifically, similar to the proof employed by Lee et al. (2015), it can be shown that if the output layer error is sufficiently small and the synaptic matrices W 1 and Y meet some conditions, then: 7/41 λU − λmaxσ(kDW 1 λC)2 2 < λU − λmaxσ(E[U (8) where kD is a conductance term and E[·] denotes the expected value. (See Theorem 1 for the proof and a concrete description of the conditions). In plain language, equation (8) says that when we consider the difference between the output target and the activity that the hidden target would have induced at the output layer, it is less than the difference between the output target and the expected value of the output layer activity during the forward phase. In other words, the output layer's error would have, on average, been smaller if the hidden layer had achieved its own target during the forward phase. Hence, if we reduce the hidden layer error, it should also reduce the output layer error, thereby providing a guarantee of appropriate credit assignment. 2 f ])2 With these error functions, we update the weights in the network with local gradient descent: ∆W 1 ∝ ∂L1 ∂W 1 ∆W 0 ∝ ∂L0 ∂W 0 (9) where ∆W i refers to the update term for weight matrix W i (see Materials and Methods, equations (19), (20), (22) and (23) for details of the weight update procedures). Given the coordination implied by equation (8), as the hidden layer reduces its own local error with gradient descent, the output layer's error should also be reduced, i.e. hidden layer learning should imply output layer learning. To test that we were successful in credit assignment for the hidden layer, and to provide empirical support for the proof, we compared the local error at the hidden layer to the output layer error across all of the image presentations to the network. We observed that, generally, whenever the hidden layer error was low, the output layer error was also low. For example, when we consider the errors for the set of '2' images presented to the network during the second epoch, there was a Pearson correlation coefficient between L0 and L1 of r = 0.61, which was much higher than what was observed for shuffled data, wherein output and hidden activities were randomly paired (Figure 4B). Furthermore, these correlations were observed across all epochs of training, with most correlation coefficients for the hidden and output errors falling between r = 0.2 - 0.6, which was, again, much higher than the correlations observed for shuffled data (Figure 4C). Interestingly, the correlations between L0 and L1 were smaller on the first epoch of training. This suggests that the coordination between the layers may only come into full effect once the network has engaged in some learning. Therefore, we inspected whether the conditions on the synaptic matrices that are assumed in the proof were, in fact, being met. More precisely, the proof assumes that the feedforward and feedback synaptic matrices (W 1 and Y , respectively) produce forward and backward transformations between the output and hidden layer that are approximate inverses of each other (see Proof of Theorem 1). Since we begin learning with random matrices, this condition is almost definitely not met at the start of training. But, we found that the network learned to meet this condition. Inspection of W 1 and Y showed that during the first epoch the forward and backwards functions became approximate inverses of each other (Figure 4, Supplement 1). This means that during the first few image presentations the network was actually learning to do credit assignment. This result is very similar to previous models examining feedback alignment (Lillicrap et al., 2016), and shows that the feedback alignment (Lillicrap et al., 2016) and difference target propagation (Lee et al., 2015) algorithms are intimately linked. Furthermore, our results suggest that very early development may involve a period of learning how to assign credit appropriately. Altogether, our model demonstrates that credit assignment using random feedback weights is a general principle that can be implemented using segregated dendrites. Deep learning with segregated dendrites Given our finding that the network was successfully assigning credit for the output error to the hidden layer neurons, we had reason to believe that our network with local weight-updates would exhibit deep 8/41 learning, i.e. an ability to take advantage of a multi-layer structure (Bengio and LeCun, 2007). To test this, we examined the effects of including hidden layers. If deep learning is indeed operational in the network, then the inclusion of hidden layers should improve the ability of the network to classify images. We built three different versions of the network (Figure 5A). The first was a network that had no hidden layer, i.e. the input neurons projected directly to the output neurons. The second was the network illustrated in Figure 3A, with a single hidden layer. The third contained two hidden layers, with the output layer projecting directly back to both hidden layers. This direct projection allowed us to build our local targets for each hidden layer using the plateaus driven by the output layer, thereby avoiding a "backward pass" through the entire network as has been used in other models (Lillicrap et al., 2016; Lee et al., 2015; Liao et al., 2015). We trained each network on the 60,000 MNIST training images for 60 epochs, and recorded the percentage of images in the 10,000 test image set that were incorrectly classified. The network with no hidden layers rapidly learned to classify the images, but it also rapidly hit an asymptote at an average error rate of 8.3% (Figure 5B, gray line). In contrast, the network with one hidden layer did not exhibit a rapid convergence to an asymptote in its error rate. Instead, it continued to improve throughout all 60 epochs, achieving an average error rate of 4.1% by the 60th epoch (Figure 5B, blue line). Similar results were obtained when we loosened the synchrony constraints and instead allowed each hidden layer neuron to engage in plateau potentials at different times (Figure 5, Supplement 1). This demonstrates that strict synchrony in the plateau potentials is not required. But, our target definitions do require two different plateau potentials separated by the teaching signal input, which mandates some temporal control of plateau potentials in the system. Interestingly, we found that the addition of a second hidden layer further improved learning. The network with two hidden layers learned more rapidly than the network with one hidden layer and achieved an average error rate of 3.2% on the test images by the 60th epoch, also without hitting a clear asymptote in learning (Figure 5B, red line). However, it should be noted that additional hidden layers beyond two did not significantly improve the error rate (data not shown), which suggests that our particular algorithm could not be used to construct very deep networks as is. Nonetheless, our network was clearly able to take advantage of multi-layer architectures to improve its learning, which is the key feature of deep learning (Bengio and LeCun, 2007; LeCun et al., 2015). Another key feature of deep learning is the ability to generate representations in the higher layers of a network that capture task-relevant information while discarding sensory details (LeCun et al., 2015; Mnih et al., 2015). To examine whether our network exhibited this type of abstraction, we used the t-Distributed Stochastic Neighbor Embedding algorithm (t-SNE). The t-SNE algorithm reduces the dimensionality of data while preserving local structure and non-linear manifolds that exist in high- dimensional space, thereby allowing accurate visualization of the structure of high-dimensional data (Maaten and Hinton, 2008). We applied t-SNE to the activity patterns at each layer of the two hidden layer network for all of the images in the test set after 60 epochs of training. At the input level, there was already some clustering of images based on their categories. However, the clusters were quite messy, with different categories showing outliers, several clusters, or merged clusters (Figure 5C, bottom). For example, the '2' digits in the input layer exhibited two distinct clusters separated by a cluster of '7's: one cluster contained '2's with a loop and one contained '2's without a loop. Similarly, there were two distinct clusters of '4's and '9's that were very close to each other, with one pair for digits on a pronounced slant and one for straight digits (Figure 5C, bottom, example images). Thus, although there is built-in structure to the categories of the MNIST dataset, there are a number of low-level features that do not respect category boundaries. In contrast, at the first hidden layer, the activity patterns were much cleaner, with far fewer outliers and split/merged clusters (Figure 5C, middle). For example, the two separate '2' digit clusters were much closer to each other and were now only separated by a very small cluster of '7's. Likewise, the '9' and '4' clusters were now distinct and no longer split based on the slant of the digit. Interestingly, when we examined the activity patterns at the second hidden layer the categories were even better segregated with only a bit of splitting or merging of category clusters (Figure 5C, top). Therefore, the network had learned to develop representations in the hidden layers wherein the categories were very distinct and low-level features unrelated to the categories were largely ignored. This abstract representation is likely to be key to the improved error rate in the two hidden layer network. 9/41 Altogether, our data demonstrates that our network with segregated dendritic compartments can engage in deep learning. Coordinated local learning mimics backpropagation of error The backpropagation of error algorithm (Rumelhart et al., 1986) is still the primary learning algorithm used for deep supervised learning in artificial neural networks (LeCun et al., 2015). Previous work has shown that learning with random feedback weights can actually match the synaptic weight updates specified by the backpropagation algorithm after a few epochs of training (Lillicrap et al., 2016). This fascinating observation suggests that deep learning with random feedback weights is not so much a different algorithm than backpropagation of error, but rather, networks with random feedback connections learn to approximate credit assignment as it is done in backpropagation (Lillicrap et al., 2016). Hence, we were curious as to whether or not our network was, in fact, learning to approximate the synaptic weight updates prescribed by backpropagation. To test this, we trained our one hidden layer network as before, but now, in addition to calculating the vector of hidden layer synaptic weight updates specified by our local learning rule (∆W 0 in equation (9)), we also calculated the vector of hidden layer synaptic weight updates that would be specified by non-locally backpropagating the error from the output layer, (∆W 0 BP ). We then calculated the angle between these two alternative weight updates. In a very high-dimensional BP ≈ 90◦). space, any two independent vectors will be roughly orthogonal to each other (i.e. ∆W 0(cid:54) ∆W 0 BP < 90◦), If the two synaptic weight update vectors are not orthogonal to each other (i.e. ∆W 0(cid:54) ∆W 0 then it suggests that the two algorithms are specifying similar weight updates. As in previous work (Lillicrap et al., 2016), we found that the initial weight updates for our network were orthogonal to the updates specified by backpropagation. But, as the network learned the angle dropped to approximately 65◦, before rising again slightly to roughly 70◦ (Figure 6A, blue line). This suggests that our network was learning to develop local weight updates in the hidden layer that were in rough agreement with the updates that explicit backpropagation would produce. However, this drop in orthogonality was still much less than that observed in non-spiking artificial neural networks learning with random feedback weights, which show a drop to below 45◦ (Lillicrap et al., 2016). We suspected that the higher angle between the weight updates that we observed may have been because we were using spikes to communicate the feedback from the upper layer, which could introduce both noise and bias in the estimates of the output layer activity. To test this, we also examined the weight updates that our algorithm would produce if we propagated the spike rates of the output layer neurons, λU (t), back directly through the random feedback weights, Y . In this scenario, we observed a much sharper drop in BP angle, which reduced to roughly 35◦ before rising again to 40◦ (Figure 6A, red line). the ∆W 0(cid:54) ∆W 0 These results show that, in principle, our algorithm is learning to approximate the backpropagation algorithm, though with some drop in accuracy introduced by the use of spikes to propagate output layer activities to the hidden layer. To further examine how our local learning algorithm compared to backpropagation we compared the low-level features that the two algorithms learned. To do this, we trained the one hidden layer network with both our algorithm and backpropagation. We then examined the receptive fields (i.e. the synaptic weights) produced by both algorithms in the hidden layer synapses (W 0) after 60 epochs of training. The two algorithms produced qualitatively similar receptive fields (Figure 6C). Both produced receptive fields with clear, high-contrast features for detecting particular strokes or shapes. To quantify the similarity, we conducted pair-wise correlation calculations for the receptive fields produced by the two algorithms and identified the maximum correlation pairs for each. Compared to shuffled versions of the receptive fields, there was a very high level of maximum correlation (Figure 6B), showing that the receptive fields were indeed quite similar. Thus, the data demonstrate that our learning algorithm using random feedback weights into segregated dendrites can in fact come to approximate the backpropagation of error algorithm. 10/41 Conditions on feedback weights Once we had convinced ourselves that our learning algorithm was, in fact, producing deep learning similar to that produced by backpropagation of error, we wanted to examine some of the constraints on learning. First, we wanted to explore the structure of the feedback weights. In our initial simulations we used non-sparse, random (i.e. normally distributed) feedback weights. We were interested in whether learning could still work with sparse weights, given that neocortical connectivity is sparse. As well, we wondered whether symmetric weights would improve learning, which would be expected given previous findings (Lillicrap et al., 2016; Lee et al., 2015; Liao et al., 2015). To explore these questions, we trained our one hidden layer network using both sparse feedback weights (only 20% non-zero values) and symmetric weights (Y = W 1T ) (Figure 7A,C). We found that learning actually improved slightly with sparse weights (Figure 7B, red line), achieving an average error rate of 3.7% by the 60th epoch, compared to the average 4.1% error rate achieved with fully random weights. But, this result appeared to depend on the magnitude of the sparse weights. To compensate for the loss of 80% of the weights we initially increased the sparse synaptic weight magnitudes by a factor of 5. However, when we did not re-scale the sparse weights learning was actually worse (Figure 7, Supplement 1). This suggests that sparse feedback provides a signal that is sufficient for credit assignment, but only if it is of appropriate magnitude. Similar to sparse feedback weights, symmetric feedback weights also improved learning, leading to a rapid decrease in the test error and an error rate of 3.6% by the 60th epoch (Figure 7D, red line). However, when we added noise to the symmetric weights this advantage was eliminated and learning was, in fact, slightly impaired (Figure 7D, blue line). At first, this was a very surprising result: given that learning works with random feedback weights, why would it not work with symmetric weights with noise? However, when we considered our previous finding that during the first epoch the feedforward weights, W 1, learn to match the inverse of the feedback function (Figure 4, Supplement 1) a possible answer becomes clear. In the case of symmetric feedback weights the synaptic matrix Y is changing as W 1 changes. This works fine when Y is set to W 1T , since that artificially forces weight alignment. But, if the feedback weights are set to W 1T plus noise, then the system can never align the weights appropriately, since Y is now a moving target. This would imply that any implementation of feedback learning must either be very effective (to achieve the right feedback) or very slow (to allow the feedforward weights to adapt). Learning with partial apical attenuation Another constraint that we wished to examine was whether total segregation of the apical inputs as we had done was necessary, given that real pyramidal neurons only show an attenuation of distal apical inputs to the soma (Larkum et al., 1999). To examine this, we re-ran our two hidden layer network, but now, we allowed the apical dendritic voltage to influence the somatic voltage by setting gA = 0.05. This value gave us twelve times more attenuation than the attenuation from the basal compartments (since gB = 0.6). This difference in the levels of attenuation is in-line with experimental data (Larkum et al., 1999, 2009) (Figure 8A). When we compared the learning in this scenario to the scenario with total apical segregation, we observed very little difference in the error rates on the test set (Figure 8B, gray and red lines). Importantly, though, we found that if we decreased the apical attenuation to the same level as the basal compartment (gA = gB = 0.6) then the learning was significantly impaired (Figure 8B, blue line). This demonstrates that although total apical attenuation is not necessary, partial segregation of the apical compartment from the soma is necessary. This result makes sense given that our local targets for the hidden layer neurons incorporate a term that is supposed to reflect the response of the output neurons to the feedforward sensory information (αf ). Without some sort of separation of feedforward and feedback information, as is assumed in other models of deep learning (Lillicrap et al., 2016; Lee et al., 2015), this feedback signal would get corrupted by recurrent dynamics in the network. Our data show that electrontonically segregated dendrites is one potential way to achieve the required separation between feedforward and feedback information. 11/41 Discussion Deep learning has radically altered the field of AI, demonstrating that parallel distributed processing across multiple layers can produce human/animal-level capabilities in image classification, pattern recognition and reinforcement learning (Hinton et al., 2006; LeCun et al., 2015; Mnih et al., 2015; Silver et al., 2016; Krizhevsky et al., 2012; He et al., 2015). Deep learning was motivated by analogies to the real brain (LeCun et al., 2015; Cox and Dean, 2014), so it is tantalizing that recent studies have shown that deep neural networks develop representations that strongly resemble the representations observed in the mammalian neocortex (Khaligh-Razavi and Kriegeskorte, 2014; Yamins and DiCarlo, 2016; Cadieu et al., 2014; Kubilius et al., 2016). In fact, deep learning models can match cortical representations even more so than some models that explicitly attempt to mimic the real brain (Khaligh-Razavi and Kriegeskorte, 2014). Hence, at a phenomenological level, it appears that deep learning, defined as multilayer cost function reduction with appropriate credit assignment, may be key to the remarkable computational prowess of the mammalian brain (Marblestone et al., 2016). However, the lack of biologically feasible mechanisms for credit assignment in deep learning algorithms, most notably backpropagation of error (Rumelhart et al., 1986), has left neuroscientists with a mystery. How can the real brain solve the credit assignment problem (Figure 1)? Here, we expanded on an idea that previous authors have explored (Kording and Konig, 2001; Spratling, 2002; Spratling and Johnson, 2006) and demonstrated that segregating the feedback and feedforward inputs to neurons, much as the real neocortex does (Larkum et al., 1999, 2007, 2009), can enable the construction of local targets to assign credit appropriately to hidden layer neurons (Figure 2). With this formulation, we showed that we could use segregated dendritic compartments to coordinate learning across layers (Figure 3 and Figure 4). This enabled our network to take advantage of multiple layers to develop representations of hand-written digits in hidden layers that enabled better levels of classification accuracy on the MNIST dataset than could be achieved with a single layer (Figure 5). Furthermore, we found that our algorithm actually approximated the weight updates that would be prescribed by backpropagation, and produced similar low-level feature detectors (Figure 6). As well, we showed that our basic framework works with sparse feedback connections (Figure 7) and more realistic, partial apical attenuation (Figure 8). Therefore, our work demonstrates that deep learning is possible in a biologically feasible framework, provided that feedforward and feedback signals are sufficiently segregated in different dendrites. Perhaps the most biologically unrealistic component of backpropagation is the use of non-local "weight-transport" (Figure 1B) (Grossberg, 1987). Our model builds on recent neural networks research demonstrating that weight transport is not, in fact, required for credit assignment (Lillicrap et al., 2016; Liao et al., 2015; Lee et al., 2015). By demonstrating that the credit assignment problem is solvable using feedback to generate spatially local weight updates, these studies have provided a major breakthrough in our understanding of how deep learning could work in the real brain. However, this previous research involved an implicit assumption of separate feedforward and feedback pathways (Figure 2A). Although this is a possibility, there is currently no evidence for separate feedback pathways to guide learning in the neocortex. In our study, we obtained separate feedforward and feedback information using electrotonically segregated dendrites (Figure 2B), in analogy to neocortical pyramidal neurons (Larkum et al., 1999). This segregation allowed us to perform credit assignment using direct feedback pathways between layers (Figures 3-5). Moreover, we found that even with partial attenuation of the conductance from the apical dendrites to the soma our network could engage in deep learning (Figure 8). It should be recognized, though, that although our learning algorithm achieved deep learning with spatially local update rules, we had to assume some temporal non-locality. Our credit assignment system depended on the difference between the target and forward plateaus, αt − αf , which occurred at different times (roughly a 50-60 ms gap). Although this is not an ideal solution, this small degree of temporal non-locality for synaptic update rules is in-line with known biological mechanisms, such as slowly decaying calcium transients or synaptic tags (Redondo and Morris, 2011). Hence, our model exhibited deep learning using only local information contained within the cells. In this work we adopted a similar strategy to the one taken by Lee et al.'s (2015) difference target propagation algorithm, wherein the feedback from higher layers is used to construct local activity targets 12/41 at the hidden layers. One of the reasons that we adopted this strategy is that it is appealing to think that feedback from upper layers may not simply be providing a signal for plasticity, but also a modulatory signal to push the hidden layer neurons towards a "better" activity pattern in real-time. This sort of top-down modulation could be used by the brain to improve sensory processing in different contexts and engage in inference (Bengio et al., 2015). Indeed, framing cortico-cortical feedback as a mechanism to modulate incoming sensory activity is a more common way of viewing feedback signals in the neocortex (Larkum, 2013; Gilbert and Li, 2013; Zhang et al., 2014; Fiser et al., 2016). In light of this, it is interesting to note that distal apical inputs in somatosensory cortex can help animals perform sensory discrimination tasks (Takahashi et al., 2016; Manita et al., 2015). However, in our model, we did not actually implement a system that altered the hidden layer activity to make sensory calculations-we simply used the feedback signals to drive learning. In-line with this view of top-down feedback, two recent papers have found evidence that cortical feedback can indeed guide feedforward sensory plasticity (Thompson et al., 2016; Yamada et al., 2017). Yet, there is no reason that feedback signals cannot provide both top-down modulation and a signal for learning (Spratling, 2002). In this respect, a potential future advance on our model would be to implement a system wherein the feedback actively "nudges" the hidden layers towards appropriate activity patterns in order to guide learning while also shaping perception. This proposal is reminiscent of the approach taken in previous computational models (Urbanczik and Senn, 2014; Spratling and Johnson, 2006; Kording and Konig, 2001). Future research could study how top-down modulation and a signal for credit assignment can be combined in deep learning models. It is important to note that there are many aspects of our model that are not biologically realistic. These include (but are not limited to) synaptic inputs that are not conductance based, synaptic weights that can switch from positive to negative (and vice-versa), and plateau potentials that are considered instantaneous and are not an accurate reflection of calcium spike dynamics in real pyramidal neurons (Larkum et al., 1999, 2009). However, the intention with this study was not to achieve total biological realism, but rather to demonstrate that the sort of dendritic segregation that neocortical pyramidal neurons exhibit can subserve credit assignment. Although we found that deep learning can still work with realistic levels of passive conductance from the apical compartment to the somatic compartment (Figure 8B, red line) (Larkum et al., 1999), we also found that learning was impaired if the basal and apical compartments had equal conductance to the soma (Figure 8B, blue line). This is interesting, because it suggests that the unique physiology of neocortical pyramidal neurons may in fact reflect nature's solution to deep learning. Perhaps the relegation of feedback from higher-order brain regions to the electrotonically distant apical dendrites (Budd, 1998; Spratling, 2002), and the presence of active calcium spike mechanisms in the apical shaft (Larkum et al., 2009), are mechanisms for coordinating local synaptic weight updates. If this is correct, then the inhibitory interneurons that target apical dendrites and limit active communication to the soma (Murayama et al., 2009) may be used by the neocortex to control learning. Although this is speculative, it is worth noting that current evidence supports the idea that neuromodulatory inputs carrying temporally precise salience information (Hangya et al., 2015) can shut off interneurons to disinhibit the distal apical dendrites (Pi et al., 2013; Karnani et al., 2016; Pfeffer et al., 2013; Brombas et al., 2014), and presumably, promote apical communication to the soma. Recent work suggests that the specific patterns of interneuron inhibition on the apical dendrites are very spatially precise and differentially timed to motor behaviours (Munoz et al., 2017), which suggests that there may well be coordinated physiological mechanisms for determining when cortico-cortical feedback is transmitted to the soma. Future research should examine whether these inhibitory and neuromodulatory mechanisms do, in fact, open the window for plasticity in the basal dendrites of pyramidal neurons, as our model, and some recent experimental work (Bittner et al., 2015), would predict. An additional issue that should be recognized is that the error rates which our network achieved were by no means as low as can be achieved with artificial neural networks, nor at human levels of performance (LeCun et al., 1998; Li et al., 2016). As well, our algorithm was not able to take advantage of very deep structures (beyond two hidden layers, the error rate did not improve). In contrast, increasing the depth of networks trained with backpropagation can lead to performance improvements (Li et al., 2016). But, these observations do not mean that our network was not engaged in deep learning. First, it is interesting to note that although the backpropagation algorithm is several decades old (Rumelhart et al., 1986), it 13/41 was long considered to be useless for training networks with more than one or two hidden layers (Bengio and LeCun, 2007). Indeed, it was only the use of layer-by-layer training that initially led to the realization that deeper networks can achieve excellent performance (Hinton et al., 2006). Since then, both the use of very large datasets (with millions of examples), and additional modifications to the backpropagation algorithm, have been key to making backpropagation work well on deeper networks (Sutskever et al., 2013; LeCun et al., 2015). Future studies could examine how our algorithm could incorporate current techniques used in machine learning to work better on deeper architectures. Second, we stress that our network was not designed to match the state-of-the-art in machine learning, nor human capabilities. To test our basic hypothesis (and to run our leaky-integration and spiking simulations in a reasonable amount of time) we kept the network small, we stopped training before it reached its asymptote, and we did not implement any add-ons to the learning to improve the error rates, such as convolution and pooling layers, initialization tricks, mini-batch training, drop-out, momentum or RMSProp (Sutskever et al., 2013; Tieleman and Hinton, 2012; Srivastava et al., 2014). Indeed, it would be quite surprising if a relatively vanilla, small network like ours could come close to matching current performance benchmarks in machine learning. Third, although our network was able to take advantage of multiple layers to improve the error rate, there may be a variety of reasons that ever increasing depth didn't improve performance significantly. For example, our use of direct connections from the output layer to the hidden layers may have impaired the network's ability to coordinate synaptic updates between hidden layers. As well, given our finding that the use of spikes produced weight updates that were less well-aligned to backpropagation (Figure 6A) it is possible that deeper architectures require mechanisms to overcome the inherent noisiness of spikes. One aspect of our model that we did not develop was the potential for learning at the feedback synapses. Although we used random synaptic weights for feedback, we also demonstrated that our model actually learns to meet the mathematical conditions required for credit assignment, namely the feedforward weights come to approximate the inverse of the feedback (Figure 4, Supplement 1). This suggests that it would be beneficial to develop a synaptic weight update rule for the feedback synapses that made this aspect of the learning better. Indeed, Lee et al. (2015) implemented an "inverse loss function" for their feedback synapses which promoted the development of feedforward and feedback functions that were roughly inverses of each other, leading to the emergence of auto-encoder functions in their network. In light of this, it is interesting to note that there is evidence for unique, "reverse" spike- timing-dependent synaptic plasticity rules in the distal apical dendrites of pyramidal neurons (Sjostrom and Hausser, 2006; Letzkus et al., 2006), which have been shown to produce symmetric feedback weights and auto-encoder functions in artificial spiking networks (Burbank and Kreiman, 2012; Burbank, 2015). Thus, it is possible that early in development the neocortex actually learns cortico-cortical feedback connections that help it to assign credit for later learning. Our work suggests that any experimental evidence showing that feedback connections learn to approximate the inverse of feedforward connections could be considered as evidence for deep learning in the neocortex. In summary, deep learning has had a huge impact on AI, but, to date, its impact on neuroscience has been limited. Nonetheless, given a number of findings in neurophysiology and modeling (Yamins and DiCarlo, 2016), there is growing interest in understanding how deep learning may actually be achieved by the real brain (Marblestone et al., 2016). Our results show that by moving away from point neurons, and shifting towards multi-compartment neurons that segregate feedforward and feedback signals, the credit assignment problem can be solved and deep learning can be achieved. Perhaps the dendritic anatomy of neocortical pyramidal neurons is important for nature's own deep learning algorithm. Materials and Methods Code for the model can be obtained from a GitHub repository (https://github.com/jordan-g/Segregated- Dendrite-Deep-Learning) (Guergiuev, 2017). For notational simplicity, we describe our model in the case of a network with only one hidden layer. We describe how this is extended to a network with multiple layers at the end of this section. As well, at the end of this section in Table 1 we provide a table listing 14/41 the parameter values we used for all of the simulations presented in this paper. Neuronal dynamics The network described here consists of an input layer with (cid:96) neurons, a hidden layer with m neurons, and an output layer with n neurons. Neurons in the input layer are simple Poisson spiking neurons whose rate-of-fire is determined by the intensity of image pixels (ranging from 0 - λmax). Neurons in the hidden layer are modeled using three functional compartments-basal dendrites with voltages B(t) = [B1(t), B2(t), ..., Bm(t)], apical dendrites with voltages A(t) = [A1(t), A2(t), ..., Am(t)], and somata with voltages C(t) = [C1(t), C2(t), ..., Cm(t)]. Feedforward inputs from the input layer and feedback inputs from the output layer arrive at basal and apical synapses, respectively. At basal synapses, presynaptic spikes are translated into post-synaptic potentials P SP B(t) = [P SP B (cid:96) (t)] given by: 2 (t), ..., P SP B 1 (t), P SP B P SP B j (t) = κ(t − s) (10) (cid:88) s∈Xj (cid:96)(cid:88) n(cid:88) j=1 where Xj is the set of pre-synaptic spike times at synapses with presynaptic input neuron j. κ(t) = (e−t/τL − e−t/τs)Θ(t)/(τL − τs) is the response kernel, where τs and τL are short and long time constants, and Θ is the Heaviside step function. The post-synaptic potentials at apical synapses, P SP A(t) = [P SP A n (t)], are modeled in the same manner. The basal and apical dendritic potentials for neuron i are then given by weighted sums of the post-synaptic potentials at either its basal or apical synapses: 2 (t), ..., P SP A 1 (t), P SP A Bi(t) = Ai(t) = W 0 ijP SP B j (t) + b0 i YijP SP A j (t) (11) where b0 = [b0 m] are bias terms, W 0 is the m × (cid:96) matrix of feedforward weights for neurons in the hidden layer, and Y is the m× n matrix of their feedback weights. The somatic voltage for neuron i evolves with leak as: 2, ..., b0 1, b0 j=1 dCi(t) dt = −gLCi(t) + gB(Bi(t) − Ci(t)) + gA(Ai(t) − Ci(t)) (12) where gL is the leak conductance, gB is the conductance from the basal dendrite to the soma, and gA is the conductance from the apical dendrite to the soma. Equation (12) is identical to equation (1) in Results. Note that for simplicity's sake we are assuming a resting potential of 0 V and a membrane capacitance of 1 F, but these values are not important for the results. The instantaneous firing rates of neurons in the hidden layer are given by λC(t) = [λC 2 (t), ..., λC m(t)], i (t) is the result of applying a nonlinearity, σ(·), to the somatic potential Ci(t). We chose σ(·) 1 (t), λC where λC to be a simple sigmoidal function, such that: λC i (t) = λmaxσ(Ci(t)) = λmax 1 1 + e−Ci(t) (13) Here, λmax is the maximum possible rate-of-fire for the neurons, which we set to 200 Hz. Note that equation (13) is identical to equation (2) in Results. Spikes are then generated using Poisson processes with these firing rates. We note that although the maximum rate was 200 Hz, the neurons rarely achieved 15/41 m(cid:88) anything close to this rate, and the average rate of fire in the neurons during our simulations was 24 Hz, in-line with neocortical firing rates (Steriade et al., 2001). Units in the output layer are modeled using only two compartments, dendrites with voltages V (t) = [V1(t), V2(t), ..., Vn(t)] and somata with voltages U (t) = [U1(t), U2(t), ..., Un(t)]. Vi(t) is given by: Vi(t) = W 1 ijP SP V j (t) + b1 i (14) j=1 where P SP V (t) = [P SP V m (t)] are the post-synaptic potentials at synapses that receive feedforward input from the hidden layer, and are calculated in the manner described by equation (10). Ui(t) evolves as: 2 (t), ..., P SP V 1 (t), P SP V dUi(t) dt = −gLUi(t) + gD(Vi(t) − Ui(t)) + Ii(t) (15) where gL is the leak conductance, gD is the conductance from the dendrite to the soma, and I(t) = [I1(t), I2(t), ..., In(t)] are somatic currents that can drive output neurons toward a desired somatic voltage. For neuron i, Ii is given by: Ii(t) = gEi(t)(EE − Ui(t)) + gIi(t)(EI − Ui(t)) (16) where gE(t) = [gE1 (t), gE2(t), ..., gEn (t)] and gI (t) = [gI1 (t), gI2(t), ..., gIn (t)] are time-varying exci- tatory & inhibitory nudging conductances, and EE and EI are the excitatory and inhibitory reversal potentials. In our simulations, we set EE = 8 V and EI = −8 V. During the target phase only, we set gIi = 1 and gEi = 0 for all units i whose output should be minimal, and gEi = 1 and gIi = 0 for the unit whose output should be maximal. In this way, all units other than the "target" unit are silenced, while the "target" unit receives a strong excitatory drive. In the forward phase, I(t) is set to 0. The Poisson spike rates λU (t) = [λU n (t)] are calculated as in equation (13). 2 (t), ..., λU 1 (t), λU Plateau potentials At the end of the forward and target phases, we calculate plateau potentials αf = [αf αt = [αt m] for apical dendrites of hidden layer neurons, where αf 2, ..., αt i and αt 1, αt 1 , αf 2 , ..., αf i are given by: m] and (cid:90) t1 (cid:90) t2 t1−tp t2−tp αf i = σ( αt i = σ( Ai(t)dt) Ai(t)dt) (17) where t1 and t2 are the end times of the forward and target phases, respectively, and tp = ti+1 − (ti + ∆ts) is the integration time for the plateau potential. ∆ts = 30 ms is the settling time for the voltages. Note that equation (17) is identical to equation (3) in Results. These plateau potentials are used by hidden layer neurons to update their basal weights. Weight updates All feedforward synaptic weights are updated at the end of each target phase. Output layer units update their synaptic weights W 1 in order to minimize the loss function L1 given in equation (7). All average voltages are calculated after a delay ∆ts from the start of a phase, which allows for the network to 16/41 reach a steady state before averaging begins. In practice this means that the average somatic voltage for output layer neuron i in the forward phase, Ui , has the property f (cid:0) m(cid:88) (cid:1) f ≈ kDVi f Ui = kD W 1 ijP SP V j f + b1 i where kD = gD/(gL + gD). Thus, j=1 ∂L1 ∂W 1 ≈ −kDλmax(λU − λmaxσ(U ∂b1 ≈ −kDλmax(λU − λmaxσ(U ∂L1 f ))σ(cid:48)(U f ) ◦ P SP V f f ))σ(cid:48)(U f ) (18) (19) The dendrites in the output layer use this approximation of the gradient in order to update their weights using gradient descent: W 1 → W 1 − η1P 1 ∂L1 ∂W 1 b1 → b1 − η1P 1 ∂L1 ∂b1 (20) where η1 is a learning rate constant, and P 1 is a scaling factor used to normalize the scale of the rate-of-fire function. In the hidden layer, basal dendrites update their synaptic weights W 0 by minimizing the loss function L0 given in equation (7). We define the target rates-of-fire λC = [λC 1 , λC 2 , ..., λC m] such that f λC i = λC i + αt i − αf i (21) 1 , αf where αf = [αf 2 , ..., αf m] and αt = [αt 1, αt 2, ..., αt m] are forward and target phase plateau potentials given in equation (17). Note that equation (21) is identical to equation (6) in Results. These hidden layer target firing rates are similar to the targets used in difference target propagation (Lee et al., 2015). Using the fact that λC i i − λmaxσ(Ci ), we can say that λC f ≈ λmaxσ(Ci i , and hence: i − αf ) ≈ αt f f ∂L0 ∂W 0 ≈ −kB(αt − αf )λmaxσ(cid:48)(C ∂b0 ≈ −kB(αt − αf )λmaxσ(cid:48)(C ∂L0 f ) ◦ P SP B f f ) where kB = gB/(gL + gB + gA). Basal weights are updated in order to descend this gradient: W 0 → W 0 − η0P 0 ∂L0 ∂W 0 b0 → b0 − η0P 0 ∂L0 ∂b0 (22) (23) Importantly, this update rule is fully local for the hidden layer neurons. It consists essentially of three terms, (1) the difference in the plateau potentials for the target and forward phases (αt − αf ), (2) the derivative of the spike rate function (λmaxσ(cid:48)(C ). All three of these terms are values that a real neuron could theoretically calculate using some combination of molecular synaptic tags, calcium currents, and back-propagating action potentials. )), and (3) the postsynaptic potentials (P SP B f f 17/41 Multiple hidden layers In order to extend our algorithm to deeper networks with multiple hidden layers, our model incorporates direct synaptic connections from the output layer to each hidden layer. Thus, each hidden layer receives feedback from the output layer through its own separate set of fixed, random weights. For example, in a network with two hidden layers, both layers receive the feedback from the output layer at their apical dendrites through backward weights Y 0 and Y 1. The local targets at each layer are then given by: t λU = λU λC1 λC0 = λC1 t = λC0 t + α1t − α1f + α0t − α0f where the superscripts 0 and 1 denote the first and second hidden layers, respectively. The local loss functions at each layer are: L2 = λU − λmaxσ(U L1 = λC1 − λmaxσ(C1 L0 = λC0 − λmaxσ(C0 )2 )2 )2 2 f f f 2 2 (24) (25) (26) (27) where L2 is the loss at the output layer. The learning rules used by the hidden layers in this scenario are the same as in the case with one hidden layer. Learning rate optimization For each of the three network sizes that we present in this paper, a grid search was performed in order to find good learning rates. We set the learning rate for each layer by stepping through the range [0.1, 0.3] with a step size of 0.02. For each combination of learning rates, a neural network was trained for one epoch on the 60, 000 training examples, after which the network was tested on 10,000 test images. The learning rates that gave the best performance on the test set after an epoch of training were used as a basis for a second grid search around these learning rates that used a smaller step size of 0.01. From this, the learning rates that gave the best test performance after 20 epochs were chosen as our learning rates for that network size. In all of our simulations, we used a learning rate of 0.19 for a network with no hidden layers, learning rates of 0.21 (output and hidden) for a network with one hidden layer, and learning rates of 0.23 (hidden layers) and 0.12 (output layer) for a network with two hidden layers. All networks with one hidden layer had 500 hidden layer neurons, and all networks with two hidden layers had 500 neurons in the first hidden layer and 100 neurons in the second hidden layer. Training paradigm For all simulations described in this paper, the neural networks were trained on classifying handwritten digits using the MNIST database of 28 pixel × 28 pixel images. Initial feedforward and feedback weights were chosen randomly from a uniform distribution over a range that was calculated to produce voltages in the dendrites between −6 - 12 V. Prior to training, we tested a network's initial performance on a set of 10,000 test examples. This set of images was shuffled at the beginning of testing, and each example was shown to the network in sequence. Each input image was encoded into Poisson spiking activity of the 784 input neurons representing each pixel of the image. The firing rate of an input neuron was proportional to the brightness of the pixel that it represents (with spike rates between [0 - λmax]. The spiking activity of each of the 18/41 784 input neurons was received by the neurons in the first hidden layer. For each test image, the network underwent only a forward phase. At the end of this phase, the network's classification of the input image was given by the neuron in the output layer with the greatest somatic potential (and therefore the greatest spike rate). The network's classification was compared to the target classification. After classifying all 10,000 testing examples, the network's classification error was given by the percentage of examples that it did not classify correctly. Following the initial test, training of the neural network was done in an on-line fashion. All 60,000 training images were randomly shuffled at the start of each training epoch. The network was then shown each training image in sequence, undergoing a forward phase ending with a plateau potential, and a target phase ending with another plateau potential. All feedforward weights were then updated at the end of the target phase. At the end of the epoch (after all 60,000 images were shown to the network), the network was again tested on the 10,000 test examples. The network was trained for up to 60 epochs. Simulation details For each training example, a minimum length of 50 ms was used for each of the forward and target phases. The lengths of the forward and target training phases were determined by adding their minimum length to an extra length term, which was chosen randomly from a Wald distribution with a mean of 2 ms and scale factor of 1. During testing, a fixed length of 500 ms was used for the forward transmit phase. Average forward and target phase voltages were calculated after a settle duration of ∆ts = 30 ms from the start of the phase. For simulations with randomly sampled plateau potential times (Figure 5, Supplement 1), the time at which each neuron's plateau potential occurred was randomly sampled from a folded normal distribution (µ = 0, σ2 = 3) that was truncated (max = 5) such that plateau potentials occurred between 0 ms and 5 ms before the start of the next phase. In this scenario, the average apical voltage in the last 30 ms was averaged in the calculation of the plateau potential for a particular neuron. The time-step used for simulations was dt = 1 ms. At each time-step, the network's state was updated bottom-to-top beginning with the first hidden layer and ending with the output layer. For each layer, dendritic potentials were updated, followed by somatic potentials, and finally their spiking activity. Table 1 lists the simulation parameters and the values that were used in the figures presented. All code was written using the Python programming language version 2.7 (RRID: SCR 008394) with the NumPy (RRID: SCR 008633) and SciPy (RRID: SCR 008058) libraries. The code is open source and is freely available at https://github.com/jordan-g/Segregated-Dendrite-Deep-Learning (Guergiuev, 2017). The data used to train the network was from the Mixed National Institute of Standards and Technology (MNIST) database, which is a modification of the original database from the National Institute of Standards and Technology (RRID: SCR 006440) (LeCun et al., 1998). The MNIST database can be found at http://yann.lecun.com/exdb/mnist/. Some of the simulations were run on the SciNet High-Performance Computing platform (Loken et al., 2010). Proofs Theorem for loss function coordination The targets that we selected for the hidden layer (see equation (6)) were based on the targets used in Lee et al. (2015). The authors of that paper provided a proof showing that their hidden layer targets guaranteed that learning in one layer helped reduce the error in the next layer. However, there were a number of differences between our network and theirs, such as the use of spiking neurons, voltages, different compartments, etc. Here, we modify the original Lee et al. (2015) proof slightly to prove Theorem 1 (which is the formal statement of equation (8)). One important thing to note is that the theorem given here utilizes a target for the hidden layer that is slightly different than the one defined in equation (6). However, the target defined in equation (6) is a 19/41 t f + σ(Y λU numerical approximation of the target given in Theorem 1. After the proof of we describe exactly how these approximations relate to the targets given here. Theorem 1. Consider a neural network with one hidden layer and an output layer. Let λC∗ λC σ(·) is a differentiable function. Assume that U be the target firing rates for the output layer. Also, for notational simplicity, let β(x) ≡ λmaxσ(kDW 1x) and γ(x) ≡ σ(Y x). Theorem 1 states that if λU − λmaxσ(E[U ]) is sufficiently small, and the Jacobian matrices Jβ and Jγ satisfy the condition that the largest eigenvalue of (I − JβJγ)T (I − JβJγ) is less than 1, then ])) be the target firing rates for neurons in the hidden layer, where ) − σ(Y λmaxσ(E[U . Let λU = λU f ≈ kDV = f f f t λU − λmaxσ(kDW 1 λC∗ )2 2 < λU − λmaxσ(E[U f ])2 2 We note again that the proof for this theorem is essentially a modification of the proof provided in f Lee et al. (2015) that incorporates our Lemma 1 to take into account the expected value of P SP V given that spikes in the network are generated with non-stationary Poisson processes. , Proof. λU − λmaxσ(kDW 1 λC∗ ) ≡ λU − β(λC∗ = λU − β(λC f ) t + γ(λU Lemma 1 shows that λmaxσ(E[U ficiently large averaging time window. Assume that λmaxσ(E[U Then, ]) = λmaxσ(E[kDW 1P SP V f f ) − γ(λmaxσ(E[U ]))) ]) ≈ λmaxσ(kDW 1λC ]) = λmaxσ(kDW 1λC f f f ) given a suf- ) ≡ β(λC f ). f λU − β(λC∗ ) = λU − β(λC f + γ(λU t ) − γ(β(λC f ))) Let e = λU t − β(λC f ). Applying Taylor's theorem, λU − β(λC∗ ) = λU − β(λC f + Jγe + o(e2)) where o(e2) is the remainder term that satisfies lime→0 o(e2)/e2 = 0. Applying Taylor's theorem again, λU − β(λC∗ ) = λU − β(λC f ) − Jβ(Jγe + o(e2)) − o((Jγe + o(e2)2) Then, λU − β(λC∗ f = λU − β(λC = (I − JβJγ)e − o(e2) ) + JβJγe − o(e2) )2 2 = ((I − JβJγ)e − o(e2))T ((I − JβJγ)e − o(e2)) = eT (I − JβJγ)T (I − JβJγ)e − o(e2)T (I − JβJγ)e − eT (I − JβJγ)T o(e2) + o(e2)T o(e2) = eT (I − JβJγ)T (I − JβJγ)e + o(e2 2) ≤ µe2 2 + o(e2 2) 20/41 where µ is the largest eigenvalue of (I−JβJγ)T (I−JβJγ). If e is sufficiently small so that o(e2 then 2)) < (1 − µ)e2 2, λU − λmaxσ(kDW 1 λC∗ )2 2 ≤ e2 2 = λU − λmaxσ(E[U f ])2 2 Note that the last step requires that µ, the largest eigenvalue of (I − JβJγ)T (I − JβJγ), is below 1. Clearly, we do not actually have any guarantee of meeting this condition. However, our results show that even though the feedback weights are random and fixed, the feedforward weights actually learn to meet this condition during the first epoch of training (Figure 4, Supplement 1). Hidden layer targets Theorem 1 shows that if we use a target λC∗ )) for the hidden layer, there is a guarantee that the hidden layer approaching this target will also push the upper layer closer to its target λU , if certain other conditions are met. Our specific choice of λC defined in the Results (equation (6)) approximates this target rate vector using variables that are accessible to the hidden layer units. ) − σ(Y λmaxσ(kDW 1λC + σ(Y λU = λC f f t Because neuronal units calculate averages after the network has reached a steady state, λmaxσ(U f f f f ] ≈ kDW 1λC ) ≈ . If we assume that f λU U and A f ≈ E[U f f ≈ Y λU ], which is true on average, then: . Using Lemma 1 and equation (4), E[U αf = σ(A f ) ≈ σ(Y λU f ) ≈ σ(Y λmaxσ(U f )) ≈ σ(Y λmaxσ(kDW 1λC f )) and: αt = σ(A t ) ≈ σ(Y λU t ) (28) (29) Therefore, λC ≈ λC∗ Thus, our hidden layer targets ensure that our model employs a learning rule similar to difference . target propagation that approximates the necessary conditions to guarantee error convergence. Lemma for firing rates Theorem 1 had to rely on the equivalence between the average spike rates of the neurons and the postsynaptic potentials. Here, we prove a lemma showing that this equivalence does indeed hold as long as the integration time is long enough relative to the synaptic time constants ts and tL. Lemma 1. Let X be a set of presynaptic spike times during the time interval ∆t = t1 − t0, distributed according to an inhomogeneous Poisson process. Let N = X denote the number of presynaptic spikes during this time window, and let si ∈ X denote the ith presynaptic spike, where 0 < i ≤ N . Finally, let λ(t) denote the time-varying presynaptic firing rate (i.e. the time-varying mean of the Poisson process), and P SP (t) be the postsynaptic potential at time t given by equation (10). Then, during the time window ∆t, as long as ∆t (cid:29) 2τ 2 s λ2/(τL − τs)2(τL + τs), Lτ 2 E[P SP (t)] ≈ λ 21/41 Proof. The average of P SP (t) over the time window ∆t is P SP = = Since Θ(t − s) = 0 for all t < s, P SP = = 1 ∆t 1 ∆t 1 ∆t 1 ∆t EX [P SP ] = EX (cid:34) 1 ∆t EX [N ] (cid:90) t1 (cid:88) t0 P SP (t)dt (cid:90) t1 s∈X t0 s s∈X (cid:90) t1 (cid:88) (cid:32) N −(cid:88) (cid:32) N −(cid:88) (cid:34) N(cid:88) s∈X s∈X − 1 ∆t e−(t−s)/τL − e−(t−s)/τs τL − τs Θ(t − s)dt e−(t−s)/τL − e−(t−s)/τs τL − τs dt τLe−(t1−s)/τL − τse−(t1−s)/τs τL − τs (cid:33) (cid:33)(cid:35) τLe−(t1−s)/τL − τse−(t1−s)/τs τL − τs (cid:32) τLe−(t1−si)/τL − τse−(t1−si)/τs (cid:33)(cid:35) The expected value of P SP with respect to X is given by g(s)fsi(s)ds 22/41 Since the presynaptic spikes are an inhomogeneous Poisson process with a rate λ, EX [N ] =(cid:82) t1 ∆t i=0 τL − τs EX = λdt. t0 Thus, (cid:90) t1 t0 1 ∆t EX [P SP ] = (cid:35) g(si) (cid:34) N(cid:88) (cid:35) i=0 EX λdt − 1 ∆t (cid:34) N(cid:88) = λ − 1 ∆t EX g(si) i=0 where we let g(si) ≡ (τLe−(t1−si)/τL − τse−(t1−si)/τs)/(τL − τs). Then, the law of total expectation gives Letting fsi (s) denote P (si = s), we have that (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)N (cid:35)(cid:35) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)N = n (cid:35) (cid:33) · P (N = n) EX g(si) = EN g(si) (cid:34) N(cid:88) (cid:35) i=0 = (cid:34) N(cid:88) i=0 EX g(si) i=0 EX (cid:34) (cid:32) (cid:34) N(cid:88) (cid:34) N(cid:88) ∞(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)N = n (cid:35) EX i=0 = n=0 g(si) n(cid:88) n(cid:88) i=0 EX [g(si)] (cid:90) t1 = i=0 t0 Then, i=0 (cid:35) g(si) = = (cid:34) N(cid:88) i=0 EX fsi(s) = = (cid:35) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)N = n λ(s) λ(t)dt (cid:82) t1 t0 λ(s) λ∆t (cid:90) t1 n(cid:88) (cid:90) t1 i=0 t0 1 λ∆t n = = λ∆t t0 (cid:34) N(cid:88) EX g(si) g(s)λ(s)ds (cid:32)(cid:90) t1 t0 (cid:32) (cid:32)(cid:90) t1 λ∆t n ∞(cid:88) n=0 1 λ∆t t0 g(s)λ(s)ds · P (N = n) g(s)λ(s)ds n · P (N = n) (cid:33) (cid:33) g(s)λ(s)ds (cid:33) (cid:33)(cid:32) ∞(cid:88) n=0 −(cid:82) t1 t0 λ(t)dt [(cid:82) t1 P (N = n) = = t0 λ(t)dt]ne n! [λ∆t]ne−(λ∆t) Now, for an inhomogeneous Poisson process with time-varying rate λ(t), Since Poisson spike times are independent, for an inhomogeneous Poisson process: for all s ∈ [t0, t1]. Since Poisson spike times are independent, this is true for all i. Thus, Thus, Then, n! (cid:32)(cid:90) t1 (cid:32)(cid:90) t1 t0 t0 (cid:33)(cid:32) ∞(cid:88) (cid:33) n=0 g(s)λ(s)ds (cid:33) n [λ∆t]n n! g(s)λ(s)ds (λ∆t)eλ∆t (cid:34) N(cid:88) i=0 EX (cid:35) g(si) = = = e−(λ∆t) λ∆t e−(λ∆t) λ∆t (cid:90) t1 g(s)λ(s)ds t0 (cid:32)(cid:90) t1 t0 (cid:33) g(s)λ(s)ds EX [P SP ] = λ − 1 ∆t The second term of this equation is always greater than or equal to 0, since g(s) ≥ 0 and λ(s) ≥ 0 for all s. Thus, EX [P SP ] ≤ λ. As well, the Cauchy-Schwarz inequality states that 23/41 (cid:90) t1 t0 g(s)λ(s)ds ≤ = (cid:115)(cid:90) t1 (cid:115)(cid:90) t1 t0 t0 (cid:115)(cid:90) t1 (cid:112) t0 g(s)2ds λ(s)2ds g(s)2ds λ2∆t (cid:90) t1 t0 g(s)2ds = (cid:90) t1 t0 (cid:16) τLe−(t1−s)/τL − τse−(t1−s)/τs (cid:17)2 ds (cid:32) 4 (cid:33) τL − τs τ 2 Lτ 2 s τL + τs 1 2(τL − τs)2 2τ 2 Lτ 2 s ≤ = (τL − τs)2(τL + τs) (cid:115) g(s)λ(s)ds ≤ (cid:90) t1 t0 (cid:112) λ2∆t 2τ 2 Lτ 2 s (τL − τs)2(τL + τs) (cid:115) 2τ 2 Lτ 2 s λ2 (τL − τs)2(τL + τs) √ ∆t = (cid:115) √ ∆t EX [P SP ] ≥ λ − 1 ∆t (cid:115) = λ − 2τ 2 Lτ 2 s λ2 (τL − τs)2(τL + τs) 2τ 2 Lτ 2 s λ2 ∆t(τL − τs)2(τL + τs) (cid:115) λ − 2τ 2 Lτ 2 s λ2 ∆t(τL − τs)2(τL + τs) ≤ EX [P SP ] ≤ λ where Thus, Therefore, Then, Thus, as long as ∆t (cid:29) 2τ 2 s λ2/(τL − τs)2(τL + τs), EX [P SP ] ≈ λ. Lτ 2 What this lemma says, effectively, is that the expected value of P SP is going to be roughly the average presynaptic rate of fire as long as the time over which the average is taken is sufficiently long in comparison to the postsynaptic time constants and the average rate-of-fire is sufficiently small. In our simulations, ∆t is always greater than or equal to 50 ms, the average rate-of-fire is approximately 20 Hz, and our time constants τL and τs are 10 ms and 3 ms, respectively. Hence, in general: 2τ 2 Lτ 2 s λ2/(τL − τs)2(τL + τs) = 2(10)2(3)2(0.02)2/(10 − 3)2(10 + 3) ≈ 0.001 (cid:28) 50 Thus, in the proof of Theorem 1, we assume EX [P SP ] = λ. 24/41 Parameter Units Value Description dt λmax τs τL ∆ts gB gA gD gL P0 P1 ms Hz ms ms ms S S S S – – 1 200 3 10 30 0.6 Time step resolution Maximum spike rate Short synaptic time constant Long synaptic time constant Settle duration for calculation of average voltages Hidden layer conductance from basal dendrite to the soma 0, 0.05, 0.6 Hidden layer conductance from apical dendrite to the soma 0.6 0.1 20/λmax 20/λ2 max Output layer conductance from dendrite to the soma Leak conductance Hidden layer error signal scaling factor Output layer error signal scaling factor Table 1. List of parameter values used in our simulations. 25/41 Acknowledgments We would like to thank Douglas Tweed, Joao Sacramento, Walter Senn, and Yoshua Bengio for helpful discussions on this work. This research was supported by two grants to B.A.R.: a Discovery Grant from the Natural Sciences and Engineering Research Council of Canada (RGPIN-2014-04947) and a Google Faculty Research Award. The authors declare no competing financial interests. Some simulations were performed on the gpc supercomputer at the SciNet HPC Consortium. SciNet is funded by: the Canada Foundation for Innovation under the auspices of Compute Canada; the Government of Ontario; Ontario Research Fund - Research Excellence; and the University of Toronto. References Bengio, Y. and LeCun, Y. (2007). Scaling learning algorithms towards AI. Large-scale kernel machines, 34(5):1–41. Bengio, Y., Lee, D.-H., Bornschein, J., and Lin, Z. (2015). Towards biologically plausible deep learning. arXiv preprint arXiv:1502.04156. Bittner, K. C., Grienberger, C., Vaidya, S. P., Milstein, A. D., Macklin, J. J., Suh, J., Tonegawa, S., and Magee, J. C. (2015). Conjunctive input processing drives feature selectivity in hippocampal CA1 neurons. Nat Neurosci, 18(8):1133–1142. Brombas, A., Fletcher, L. N., and Williams, S. R. (2014). Activity-Dependent Modulation of Layer 1 Inhibitory Neocortical Circuits by Acetylcholine. The Journal of Neuroscience, 34(5):1932. Budd, J. M. (1998). Extrastriate feedback to primary visual cortex in primates: a quantitative analysis of connectivity. Proceedings of the Royal Society of London B: Biological Sciences, 265(1400):1037–1044. Burbank, K. S. (2015). Mirrored STDP Implements Autoencoder Learning in a Network of Spiking Neurons. PLoS Comput Biol, 11(12):e1004566. Burbank, K. S. and Kreiman, G. (2012). Depression-Biased Reverse Plasticity Rule Is Required for Stable Learning at Top-Down Connections. PLoS Comput Biol, 8(3):e1002393. Buzs´aki, G. and Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science, 304(5679):1926– 1929. Cadieu, C. F., Hong, H., Yamins, D. L. K., Pinto, N., Ardila, D., Solomon, E. A., Majaj, N. J., and DiCarlo, J. J. (2014). Deep neural networks rival the representation of primate IT cortex for core visual object recognition. PLoS Comput Biol, 10(12):e1003963. Cox, D. and Dean, T. (2014). Neural Networks and Neuroscience-Inspired Computer Vision. Current Biology, 24(18):R921–R929. Crick, F. (1989). The recent excitement about neural networks. Nature, 337(6203):129–132. Dan, Y. and Poo, M.-M. (2004). Spike timing-dependent plasticity of neural circuits. Neuron, 44(1):23–30. Fiser, A., Mahringer, D., Oyibo, H. K., Petersen, A. V., Leinweber, M., and Keller, G. B. (2016). Experience-dependent spatial expectations in mouse visual cortex. Nat Neurosci, 19(12):1658–1664. Gadagkar, V., Puzerey, P. A., Chen, R., Baird-Daniel, E., Farhang, A. R., and Goldberg, J. H. (2016). Dopamine neurons encode performance error in singing birds. Science, 354(6317):1278. Gilbert, C. D. and Li, W. (2013). Top-down influences on visual processing. Nat Rev Neurosci, 14(5):350–363. 26/41 Grossberg, S. (1987). Competitive learning: From interactive activation to adaptive resonance. Cognitive science, 11(1):23–63. Guergiuev, J. (2017). Segregated-dendrite-deep-learning. Github. https://github.com/jordan-g/ Segregated-Dendrite-Deep-Learning. 23f2c66. Hangya, B., Ranade, S., Lorenc, M., and Kepecs, A. (2015). Central cholinergic neurons are rapidly recruited by reinforcement feedback. Cell, 162(5):1155–1168. Harris, K. D. (2008). Stability of the fittest: organizing learning through retroaxonal signals. Trends in Neurosciences, 31(3):130–136. He, K., Zhang, X., Ren, S., and Sun, J. (2015). Delving deep into rectifiers: Surpassing human-level performance on imagenet classification. In Proceedings of the IEEE International Conference on Computer Vision, pages 1026–1034. Hilscher, M. M., Leao, R. N., Edwards, S. J., Leao, K. E., and Kullander, K. (2017). Chrna2- Martinotti Cells Synchronize Layer 5 Type A Pyramidal Cells via Rebound Excitation. PLOS Biology, 15(2):e2001392. Hinton, G. E., Osindero, S., and Teh, Y.-W. (2006). A fast learning algorithm for deep belief nets. Neural Computation, 18(7):1527–1554. Karnani, M. M., Jackson, J., Ayzenshtat, I., Hamzehei Sichani, A., Manoocheri, K., Kim, S., and Yuste, R. (2016). Opening holes in the blanket of inhibition: Localized lateral disinhibition by VIP interneurons. The Journal of Neuroscience, 36(12):3471–3480. Khaligh-Razavi, S.-M. and Kriegeskorte, N. (2014). Deep supervised, but not unsupervised, models may explain IT cortical representation. PLoS Comput Biol, 10(11):e1003915. Krizhevsky, A., Sutskever, I., and Hinton, G. E. (2012). Imagenet classification with deep convolutional neural networks. In Advances in neural information processing systems, pages 1097–1105. Kubilius, J., Bracci, S., and Op de Beeck, H. P. (2016). Deep neural networks as a computational model for human shape sensitivity. PLoS Comput Biol, 12(4):e1004896. Kording, K. P. and Konig, P. (2001). Supervised and Unsupervised Learning with Two Sites of Synaptic Integration. Journal of Computational Neuroscience, 11(3):207–215. Larkum, M. (2013). A cellular mechanism for cortical associations: an organizing principle for the cerebral cortex. Trends in Neurosciences, 36(3):141–151. Larkum, M. E., Nevian, T., Sandler, M., Polsky, A., and Schiller, J. (2009). Synaptic Integration in Tuft Dendrites of Layer 5 Pyramidal Neurons: A New Unifying Principle. Science, 325(5941):756–760. Larkum, M. E., Waters, J., Sakmann, B., and Helmchen, F. (2007). Dendritic spikes in apical dendrites of neocortical layer 2/3 pyramidal neurons. The Journal of Neuroscience, 27(34):8999–9008. Larkum, M. E., Zhu, J. J., and Sakmann, B. (1999). A new cellular mechanism for coupling inputs arriving at different cortical layers. Nature, 398(6725):338–341. LeCun, Y., Bengio, Y., and Hinton, G. (2015). Deep learning. Nature, 521(7553):436–444. LeCun, Y., Bottou, L., Bengio, Y., and Haffner, P. (1998). Gradient-based learning applied to document recognition. Proceedings of the IEEE, 86(11):2278–2324. Lee, D.-H., Zhang, S., Fischer, A., and Bengio, Y. (2015). Difference target propagation. In Joint European Conference on Machine Learning and Knowledge Discovery in Databases, pages 498–515. Springer. 27/41 Leibo, J. Z., Liao, Q., Anselmi, F., Freiwald, W. A., and Poggio, T. (2017). View-Tolerant Face Recognition and Hebbian Learning Imply Mirror-Symmetric Neural Tuning to Head Orientation. Current Biology, 27(1):62–67. Letzkus, J. J., Kampa, B. M., and Stuart, G. J. (2006). Learning rules for spike timing-dependent plasticity depend on dendritic synapse location. Journal of Neuroscience, 26(41):10420–10429. Li, Y., Li, H., Xu, Y., Wang, J., and Zhang, Y. (2016). Very deep neural network for handwritten digit recognition. In International Conference on Intelligent Data Engineering and Automated Learning, pages 174–182. Springer. Liao, Q., Leibo, J. Z., and Poggio, T. (2015). How Important is Weight Symmetry in Backpropagation? arXiv preprint arXiv:1510.05067. Lillicrap, T. P., Cownden, D., Tweed, D. B., and Akerman, C. J. (2016). Random synaptic feedback weights support error backpropagation for deep learning. Nature Communications, 7:13276. Loken, C., Gruner, D., Groer, L., Peltier, R., Bunn, N., Craig, M., Henriques, T., Dempsey, J., Yu, C.-H., Chen, J., Dursi, L. J., Chong, J., Northrup, S., Pinto, J., Knecht, N., and Zon, R. V. (2010). Scinet: Lessons learned from building a power-efficient top-20 system and data centre. Journal of Physics: Conference Series, 256(1):012026. Maaten, L. v. d. and Hinton, G. (2008). Visualizing data using t-SNE. Journal of Machine Learning Research, 9(Nov):2579–2605. Malenka, R. C. and Bear, M. F. (2004). LTP and LTD. Neuron, 44(1):5–21. Manita, S., Suzuki, T., Homma, C., Matsumoto, T., Odagawa, M., Yamada, K., Ota, K., Matsubara, C., Inutsuka, A., Sato, M., Ohkura, M., Yamanaka, A., Yanagawa, Y., Nakai, J., Hayashi, Y., Larkum, M., and Murayama, M. (2015). A top-down cortical circuit for accurate sensory perception. Neuron, 86(5):1304–1316. Marblestone, A., Wayne, G., and Kording, K. (2016). Towards an integration of deep learning and neuroscience. arXiv preprint arXiv:1606.03813. Martin, S. J., Grimwood, P. D., and Morris, R. G. M. (2000). Synaptic plasticity and memory: an evaluation of the hypothesis. Annual Review of Neuroscience, 23(1):649–711. Mnih, V., Kavukcuoglu, K., Silver, D., Rusu, A. A., Veness, J., Bellemare, M. G., Graves, A., Riedmiller, M., Fidjeland, A. K., Ostrovski, G., Petersen, S., Beattie, C., Sadik, A., Antonoglou, I., King, H., Kumaran, D., Wierstra, D., Legg, S., and Hassabis, D. (2015). Human-level control through deep reinforcement learning. Nature, 518(7540):529–533. Murayama, M., Perez-Garci, E., Nevian, T., Bock, T., Senn, W., and Larkum, M. E. (2009). Dendritic encoding of sensory stimuli controlled by deep cortical interneurons. Nature, 457(7233):1137–1141. Munoz, W., Tremblay, R., Levenstein, D., and Rudy, B. (2017). Layer-specific modulation of neocortical dendritic inhibition during active wakefulness. Science, 355(6328):954. Pfeffer, C. K., Xue, M., He, M., Huang, Z. J., and Scanziani, M. (2013). Inhibition of inhibition in visual cortex: the logic of connections between molecularly distinct interneurons. Nat Neurosci, 16(8):1068–1076. Pi, H.-J., Hangya, B., Kvitsiani, D., Sanders, J. I., Huang, Z. J., and Kepecs, A. (2013). Cortical interneurons that specialize in disinhibitory control. Nature, 503(7477):521–524. Redondo, R. L. and Morris, R. G. M. (2011). Making memories last: the synaptic tagging and capture hypothesis. Nat Rev Neurosci, 12(1):17–30. 28/41 Rumelhart, D. E., Hinton, G. E., and Williams, R. J. (1986). Learning representations by back-propagating errors. Nature, 323:9. Scellier, B. and Bengio, Y. (2016). Towards a biologically plausible backprop. arXiv preprint arXiv:1602.05179. Schiller, J., Major, G., Koester, H. J., and Schiller, Y. (2000). NMDA spikes in basal dendrites of cortical pyramidal neurons. Nature, 404(6775):285–289. Silver, D., Huang, A., Maddison, C. J., Guez, A., Sifre, L., van den Driessche, G., Schrittwieser, J., Antonoglou, I., Panneershelvam, V., Lanctot, M., Dieleman, S., Grewe, D., Nham, J., Kalchbrenner, N., Sutskever, I., Lillicrap, T., Leach, M., Kavukcuoglu, K., Graepel, T., and Hassabis, D. (2016). Mastering the game of Go with deep neural networks and tree search. Nature, 529(7587):484–489. Sjostrom, P. J. and Hausser, M. (2006). A Cooperative Switch Determines the Sign of Synaptic Plasticity in Distal Dendrites of Neocortical Pyramidal Neurons. Neuron, 51(2):227–238. Spratling, M. W. (2002). Cortical Region Interactions and the Functional Role of Apical Dendrites. Behavioral and Cognitive Neuroscience Reviews, 1(3):219–228. Spratling, M. W. and Johnson, M. H. (2006). A feedback model of perceptual learning and categorization. Visual Cognition, 13(2):129–165. Srivastava, N., Hinton, G., Krizhevsky, A., Sutskever, I., and Salakhutdinov, R. (2014). Dropout: A simple way to prevent neural networks from overfitting. The Journal of Machine Learning Research, 15(1):1929–1958. Steriade, M., Timofeev, I., and Grenier, F. (2001). Natural Waking and Sleep States: A View From Inside Neocortical Neurons. Journal of Neurophysiology, 85(5):1969–1985. Sutskever, I., Martens, J., Dahl, G. E., and Hinton, G. E. (2013). On the importance of initialization and momentum in deep learning. ICML (3), 28:1139–1147. Takahashi, N., Oertner, T. G., Hegemann, P., and Larkum, M. E. (2016). Active cortical dendrites modulate perception. Science, 354(6319):1587. Thompson, A., Picard, N., Min, L., Fagiolini, M., and Chen, C. (2016). Cortical Feedback Regulates Feedforward Retinogeniculate Refinement. Neuron, 91(5):1021–1033. Tieleman, T. and Hinton, G. (2012). Lecture 6.5-rmsprop: Divide the gradient by a running average of its recent magnitude. COURSERA: Neural Networks for Machine Learning, 4(2). Urbanczik, R. and Senn, W. (2009). Reinforcement learning in populations of spiking neurons. Nature Neuroscience, 12(3):250. Urbanczik, R. and Senn, W. (2014). Learning by the dendritic prediction of somatic spiking. Neuron, 81(3):521–528. Yamada, Y., Bhaukaurally, K., Madar´asz, T. J., Pouget, A., Rodriguez, I., and Carleton, A. (2017). Context- and Output Layer-Dependent Long-Term Ensemble Plasticity in a Sensory Circuit. Neuron, 93(5):1198–1212.e5. Yamins, D. L. K. and DiCarlo, J. J. (2016). Using goal-driven deep learning models to understand sensory cortex. Nat Neurosci, 19(3):356–365. Zhang, S., Xu, M., Kamigaki, T., Hoang Do, J. P., Chang, W.-C., Jenvay, S., Miyamichi, K., Luo, L., and Dan, Y. (2014). Long-range and local circuits for top-down modulation of visual cortex processing. Science, 345(6197):660–665. 29/41 Zylberberg, J., Murphy, J. T., and DeWeese, M. R. (2011). A Sparse Coding Model with Synaptically Local Plasticity and Spiking Neurons Can Account for the Diverse Shapes of V1 Simple Cell Receptive Fields. PLOS Computational Biology, 7(10):e1002250. 30/41 Figures Figure 1. The credit assignment problem in multi-layer neural networks. (A) Illustration of the credit assignment problem. In order to take full advantage of the multi-circuit architecture of the neocortex when learning, synapses in earlier processing stages (blue connections) must somehow receive "credit" for their impact on behavior or cognition. However, the credit due to any given synapse early in a processing pathway depends on the downstream synaptic connections that link the early pathway to later computations (red connections). (B) Illustration of weight transport in backpropagation. To solve the credit assignment problem, the backpropagation of error algorithm explicitly calculates the credit due to each synapse in the hidden layer by using the downstream synaptic weights when calculating the hidden layer weight changes. This solution works well in AI applications, but is unlikely to occur in the real brain. 31/41 Figure 2. Potential solutions to credit assignment using top-down feedback. (A) Illustration of the implicit feedback pathway used in previous models of deep learning. In order to assign credit, feedforward information must be integrated separately from any feedback signals used to calculate error for synaptic updates (the error is indicated here with δ). (B) Illustration of the segregated dendrites proposal. Rather than using a separate pathway to calculate error based on feedback, segregated dendritic compartments could receive feedback and calculate the error signals locally. 32/41 Figure 3. Illustration of a multi-compartment neural network model for deep learning. (A) Left: Diagram of our network architecture. An input image is represented by the spiking of input units which propagate to the hidden layer through weights W 0. Hidden layer activities arrive at the output layer through weights W 1. Feedback from the output layer is sent back to the hidden layer through fixed, random weights Y . Right: Illustration of unit i in the hidden layer. The unit has a basal dendrite, soma, and apical dendrite with membrane potentials Bi, Ci and Ai, respectively. The apical dendrite communicates to the soma predominantly using non-linear signals σ(·). The spiking output of the cell, Si(t), is a Poisson process with rate λmaxσ(Ci(t)). (B) Illustration of neuronal dynamics in the two modes of processing. In the transmit mode, the apical dendrite accumulates a measure of its average membrane potential from t1 − tp to t1. In the plateau mode, an apical plateau potential equal to the nonlinearity σ(·) applied to the average apical potential travels to the soma and basal dendrite. (C) Illustration of the sequence of network phases that occur for each training example. The network undergoes a forward phase (transmit & plateau) and a target phase (transmit & plateau) in sequence. In the target phase, a teaching signal is introduced to the output layer, shaping its activity. (D) Illustration of phase length sampling. For each training example, the lengths of forward & target phases are randomly drawn from a shifted inverse Gaussian distribution with a minimum of 50 ms. 33/41 Figure 4. Co-ordinated errors between the output and hidden layers. (A) Illustration of output error (L1) and local hidden error (L0). For a given test example shown to the network in a forward phase, the output layer error is defined as the squared norm of the difference between target firing rates λU and the sigmoid function applied to the average somatic potentials U across all output units. Hidden layer error is defined similarly, except the target is λC (as defined in the text). (B) Plot of L1 vs. L0 for all of the '2' images after one epoch of training. There is a strong correlation between hidden layer error and output layer error (real data, black), as opposed to when output and hidden errors were randomly paired (shuffled data, gray). (C) Plot of correlation between hidden layer error and output layer error across training for each category of images (each dot represents one category). The correlation is significantly higher in the real data than the shuffled data throughout training. Note also that the correlation is much lower on the first epoch of training (red oval), suggesting that the conditions for credit assignment are still developing during the first epoch. f 34/41 Figure 5. Improvement of learning with hidden layers. (A) Illustration of the three networks used in the simulations. Top: a shallow network with only an input layer and an output layer. Middle: a network with one hidden layer. Bottom: a network with two hidden layers. Both hidden layers receive feedback from the output layer, but through separate synaptic connections with random weights Y 0 and Y 1. (B) Plot of test error (measured on 10,000 MNIST images not used for training) across 60 epochs of training, for all three networks described in A. The networks with hidden layers exhibit deep learning, because hidden layers decrease the test error. Right: Spreads (min – max) of the results of repeated weight tests (n = 20) after 60 epochs for each of the networks. Percentages indicate means (two-tailed t-test, 1-layer vs. 2-layer: t38 = 197.11, P38 = 2.5 × 10−58; 1-layer vs. 3-layer: t38 = 238.26, P38 = 1.9 × 10−61; 2-layer vs. 3-layer: t38 = 42.99, P38 = 2.3 × 10−33, Bonferroni correction for multiple comparisons). (C) Results of t-SNE dimensionality reduction applied to the activity patterns of the first three layers of a two hidden layer network (after 60 epochs of training). Each data point corresponds to a test image shown to the network. Points are color-coded according to the digit they represent. Moving up through the network, images from identical categories are clustered closer together and separated from images of different catefories. Thus the hidden layers learn increasingly abstract representations of digit categories. 35/41 Figure 6. Approximation of backpropagation with local learning rules. (A) Plot of the angle between weight updates prescribed by our local update learning algorithm compared to those prescribed by backpropagation of error, for a one hidden layer network over 10 epochs of training (each point on the horizontal axis corresponds to one image presentation). Data was time-averaged using a sliding window of 100 image presentations. When training the network using the local update learning algorithm, feedback was sent to the hidden layer either using spiking activity from the output layer units (blue) or by directly sending the spike rates of output units (red). The angle between the local update ∆W 0 and backpropagation weight BP remains under 90◦ during training, indicating that both algorithms point weight updates in updates ∆W 0 a similar direction. (B) Examples of hidden layer receptive fields (synaptic weights) obtained by training the network in A using our local update learning rule (left) and backpropagation of error (right) for 60 epochs. (C) Plot of correlation between local update receptive fields and backpropagation receptive fields. For each of the receptive fields produced by local update, we plot the maximum Pearson correlation coefficient between it and all 500 receptive fields learned using backpropagation (Regular). Overall, the maximum correlation coefficients are greater than those obtained after shuffling all of the values of the local update receptive fields (Shuffled). 36/41 Figure 7. Conditions on feedback synapses for effective learning. (A) Diagram of a one hidden layer network trained in B, with 80% of feedback weights set to zero. The remaining feedback weights Y (cid:48) were multiplied by 5 in order to maintain a similar overall magnitude of feedback signals. (B) Plot of test error across 60 epochs for our standard one hidden layer network (gray) and a network with sparse feedback weights (red). Sparse feedback weights resulted in improved learning performance compared to fully connected feedback weights. Right: Spreads (min – max) of the results of repeated weight tests (n = 20) after 60 epochs for each of the networks. Percentages indicate mean final test errors for each network (two-tailed t-test, regular vs. sparse: t38 = 16.43, P38 = 7.4 × 10−19). (C) Diagram of a one hidden layer network trained in D, with feedback weights that are symmetric to feedforward weights W 1, and symmetric but with added noise. Noise added to feedback weights is drawn from a normal distribution with variance σ = 0.05. (D) Plot of test error across 60 epochs of our standard one hidden layer network (gray), a network with symmetric weights (red), and a network with symmetric weights with added noise (blue). Symmetric weights result in improved learning performance compared to random feedback weights, but adding noise to symmetric weights results in impaired learning. Right: Spreads (min – max) of the results of repeated weight tests (n = 20) after 60 epochs for each of the networks. Percentages indicate means (two-tailed t-test, random vs. symmetric: t38 = 18.46, P38 = 4.3 × 10−20; random vs. symmetric with noise: t38 = −71.54, P38 = 1.2 × 10−41; symmetric vs. symmetric with noise: t38 = −80.35, P38 = 1.5 × 10−43, Bonferroni correction for multiple comparisons). 37/41 Figure 8. Importance of dendritic segregation for deep learning. (A) Left: Diagram of a hidden layer neuron with an unsegregated apical dendrite that affects the somatic potential Ci. gA represents the strength of the coupling between the apical dendrite and soma. Right: Example traces of the apical voltage Ai and the somatic voltage Ci in response to spikes arriving at apical synapses. Here gA = 0.05, so the apical activity is strongly attenuated at the soma. (B) Plot of test error across 60 epochs of training on MNIST of a two hidden layer network, with total apical segregation (gray), strong apical attenuation (red) and weak apical attenuation (blue). Apical input to the soma did not prevent learning if it was strongly attenuated, but weak apical attenuation impaired deep learning. Right: Spreads (min – max) of the results of repeated weight tests (n = 20) after 60 epochs for each of the networks. Percentages indicate means (two-tailed t-test, total segregation vs. strong attenuation: t38 = −4.00, P38 = 8.4 × 10−4; total segregation vs. weak attenuation: t38 = −95.24, P38 = 2.4 × 10−46; strong attenuation vs. weak attenuation: t38 = −92.51, P38 = 7.1 × 10−46, Bonferroni correction for multiple comparisons). 38/41 Figure Supplements Figure 4, Supplement 1. Weight alignment during first epoch of training. (A) Plot of the maximum eigenvalue of (I − JβJγ)T (I − JβJγ) over 60,000 training examples for a one hidden layer network, where Jβ and Jγ are the mean feedforward and feedback Jacobian matrices for the last 100 training examples. The maximum eigenvalue of (I − JβJγ)T (I − JβJγ) drops below 1 as learning progresses, satisfying the main condition for the learning guarantee described in Theorem 1 to hold. (B) The product of the mean feedforward and feedback Jacobian matrices, JβJγ, for a one hidden layer network, before training (left) and after 1 epoch of training (right). As training progresses, the network updates its weights in a way that causes this product to approach the identity matrix, meaning that the two matrices are roughly inverses of each other. 39/41 Figure 5, Supplement 1. Learning with stochastic plateau times. (A) Left: Raster plot showing plateau potential times during presentation of two training examples for 100 neurons in the hidden layer of a network where plateau potential times were randomly sampled for each neuron from a folded normal distribution (µ = 0, σ2 = 3) that was truncated (max = 5) such that plateau potentials occurred between 0 ms and 5 ms before the start of the next phase. In this scenario, the apical potential over the last 30 ms was integrated to calculate the plateau potential for each neuron. (B) Plot of test error across 60 epochs of training on MNIST of a one hidden layer network, with synchronized plateau potentials (gray) and with stochastic plateau potentials (red). Allowing neurons to undergo plateau potentials in a stochastic manner did not hinder training performance. 40/41 Figure 7, Supplement 1. Importance of weight magnitudes for learning with sparse weights. Plot of test error across 20 epochs of training on MNIST of a one hidden layer network, with regular feedback weights (gray), sparse feedback weights that were amplified (red), and sparse feedback weights that were not amplified (blue). The network with amplified sparse feedback weights is the same as in Figure 7A & B, where feedback weights were multiplied by a factor of 5. While sparse feedback weights that were amplified led to improved training performance, sparse weights without amplification impaired the network's learning ability. Right: Spreads (min – max) of the results of repeated weight tests (n = 20) after 20 epochs for each of the networks. Percentages indicate means (two-tailed t-test, regular vs. sparse, amplified: t38 = 44.96, P38 = 4.4 × 10−34; regular vs. sparse, not amplified: t38 = −51.30, P38 = 3.2 × 10−36; sparse, amplified vs. sparse, not amplified: t38 = −100.73, P38 = 2.8 × 10−47, Bonferroni correction for multiple comparisons). 41/41
1801.05017
1
1801
2018-01-15T21:17:37
Hypergraph based Subnetwork Extraction using Fusion of Task and Rest Functional Connectivity
[ "q-bio.NC" ]
Functional subnetwork extraction is commonly used to explore the brain's modular structure. However, reliable subnetwork extraction from functional magnetic resonance imaging (fMRI) data remains challenging due to the pronounced noise in neuroimaging data. In this paper, we proposed a high order relation informed approach based on hypergraph to combine the information from multi-task data and resting state data to improve subnetwork extraction. Our assumption is that task data can be beneficial for the subnetwork extraction process, since the repeatedly activated nodes involved in diverse tasks might be the canonical network components which comprise pre-existing repertoires of resting state subnetworks. Our proposed high order relation informed subnetwork extraction based on a strength information embedded hypergraph, (1) facilitates the multisource integration for subnetwork extraction, (2) utilizes information on relationships and changes between the nodes across different tasks, and (3) enables the study on higher order relations among brain network nodes. On real data, we demonstrated that fusing task activation, task-induced connectivity and resting state functional connectivity based on hypergraphs improves subnetwork extraction compared to employing a single source from either rest or task data in terms of subnetwork modularity measure, inter-subject reproducibility, along with more biologically meaningful subnetwork assignments.
q-bio.NC
q-bio
Hypergraph based Subnetwork Extraction using Fusion of Task and Rest Functional Connectivity Chendi Wang, Rafeef Abugharbieh Biomedical Signal and Image Computing Lab, UBC, Canada [email protected], [email protected] Abstract Functional subnetwork extraction is commonly used to explore the brain's modular structure. However, reliable subnetwork extraction from functional magnetic resonance imaging (fMRI) data remains challenging due to the pro- nounced noise in neuroimaging data. In this paper, we proposed a high order relation informed approach based on hypergraph to combine the information from multi-task data and resting state data to improve subnetwork extrac- tion. Our assumption is that task data can be beneficial for the subnetwork extraction process, since the repeatedly activated nodes involved in diverse tasks might be the canonical network components which comprise pre-existing repertoires of resting state subnetworks [1]. Our proposed high order relation informed subnetwork extraction based on a strength information embedded hypergraph, (1) facilitates the multisource integration for subnetwork extrac- tion, (2) utilizes information on relationships and changes between the nodes across different tasks, and (3) enables the study on higher order relations among brain network nodes. On real data, we demonstrated that fusing task activation, task-induced connectivity and resting state functional connectivity based on hypergraphs improves subnetwork extraction compared to employing a single source from either rest or task data in terms of subnetwork modularity measure, inter-subject reproducibility, along with more biologically meaning- ful subnetwork assignments. Keywords: Brain Subnetwork Extraction, Multisource Fusion, Functional Connectivity, Hypergraph 1 Introduction The human brain can be regarded as being a network where units, or nodes, represent different specialized regions, and edges represent communication pathways. Brain 1 network analysis methods for connectome studies include an important branch of brain subnetwork identification. Given brain connectivity matrices, brain networks can be quantitatively examined for certain commonly used network measures. The modular structure (community structure) is of particular interest; it is from this structure that we can infer information about brain subnetworks. The modular structure is extracted by subdividing a network into groups of nodes with the maxi- mal possible within-group links and minimal between-group links using community detection methods [2]. Most existing functional subnetwork extraction methods focus on resting state function connectivity data [3, 4], using functional homogeneity clustering, Independent Component Analysis (ICA), or graph community detection. However, resting state functional connectivity is inherently with low signal-to-noise ratio (SNR) and prone to false positive correlations [5]. Such noisy resting state functional connectivity in- formation leads to unreliable subnetwork extraction results. Given the resemblance between resting state and task functional subnetworks [6] and high order nodal re- lations reflected from multi-task data, we here aim to incorporate information from task data into the subnetwork extraction based on multilayer network. We explore if this integration can improve the subnetwork extraction by exploiting the mechanism of how groups of nodes collaborate together to execute a function and how these groups communicate with each other. 1.1 Related Work - Relationship between Task and Resting Functional Connectivity Recent studies indicate that resting state functional activity actually persists dur- ing task performance [7], and similar network architecture is present across task and rest, which is supported by the existence of similar multi-task Functional Connec- tivity (FC) and resting-state FC matrices that were averaged across subjects [8]. Studies have also shown that there is a strong resemblance between rest and task subnetworks [6, 9]. The spatial overlap between resting-state functional subnetworks and task-evoked activities has been discovered [10, 11]. Based on the close relationship between the two, resting state data have been used to predict the task activities, by using group ICA to discover repertories of canonical network components that will be recruited in tasks [1]; by applying the graphical connectional topology of brain regions at rest to predict functional activity of them during task [11]; or based on a voxel-matched regression method to estimate the magnitude of task-induced activity [12]. On the other hand, aggregating brain imaging data from thousands of task re- lated studies allowed the construction of 'co-activation networks', whose major com- 2 ponents and overall network topology strongly resembled functional subnetworks derived from resting-state recordings [13, 14, 15]. It has been suggested that networks involved in cognition are a subset of net- works embedded in spontaneous activity [6, 16], and a number of canonical network components in the pre-existing repertoires of intrinsic subnetworks are selectively and dynamically recruited for various cognitions [17, 1]. 1.2 Related Work - Multilayer Brain Network Analysis Multilayer network has recently been used to model and analyze complex high or- der data, such as multivariate and multiscale information within the human brain [18]. Different layers can represent relationships across different temporal variations [19], reflect different imaging modalities (such as task and rest) [18], or different frequency bands [20], etc. Hypergraph is a type of multilayer graphs, in which edges can link any number of nodes [21]. Hypergraphs have been used to identify non-random structure in structural connectivity of the cortical microcircuits [22], identify high order brain connectome biomarkers for disease prediction [23], and study relationships between functional and structural connectome data [24]. 2 High Order Relation Informed Subnetwork Ex- traction Our assumption is that task data can be beneficial for subnetwork extraction since the repeatedly activated nodes in different tasks could be the canonical network components in the spontaneous resting state subnetworks. At the same time, the multilayer structure of repeatedly activated nodes across multi-task can be elegantly presented as a hypergraph. We propose a high order relation informed subnetwork extraction model, which (1) facilitates multisource integration of task and rest data for subnetwork extraction, (2) utilizes information from the relationship between groups of activated nodes across different tasks, and (3) enables the study on higher order relations among brain network nodes. 2.1 Framework We propose a high order relation informed approach based on hypergraph to in- tegrate both resting state and task information for brain subnetwork extraction. We firstly construct a brain graph based on a certain parcellation atlas. Secondly, we detect activation of brain nodes from task data to define the nodes for mul- tiple layers in the hypergraph, and define the connection strength between nodes 3 using task-induced connectivity. Thirdly, we construct the multitask hypergraph and incorporate resting state FC strength information when setting the weights of hyperedges. Fourthly, we fuse task and rest FC using weighted combination model before performing graphcut on the constructed graph. 2.2 Notation Overview of Hypergraph 2.2.1 Notations We here follow most of the notations presented in [21]. Let V denote a set of nodes, and E denote a family of subsets e of V such that ∪e ∈ E = V . Then we define G = (V ; E) a hypergraph with the vertex set V and the hyperedge set E. A hyperedge containing just two nodes is a simple graph edge. A hyperedge e is said to be incident with a node v when v ∈ e. Two nodes are connected if they both belong to the same hyperedge. Two hyperedges are connected if the intersection of them is not an empty set, ei ∩ ej (cid:54)= ∅. Given an arbitrary set X, let X denote the cardinality of X. A hypergraph G can be represented by a V ×E incidence matrix H with entries h(v, e) = 1 if v ∈ e and 0 otherwise, see an example in Figure 1. A weighted hypergraph, G = (V ; E; w), is a hypergraph that has a positive number w(e) associated with each hyperedge e, called the weight of hyperedge e. Next, we define four important measures of hypergraph properties. the number of nodes that exist in the hyperedge. 2. We further define the hyperdegree of a hyperedge as the number of hyperedges hyperedge degrees. Take Figure 1 as an example, δ(e1) = 3, and δ(e2) = 2. For a hyperedge e ∈ E: 1. We follow [21] to define its degree as d h(e) = δ(e) := e, which counts If one uses the incidence ma- trix, δ(e) := (cid:80){v∈V } h(v, e). Let De denote the diagonal matrices containing the connected to it, denoted as d hH(e) :=(cid:80){ei∈E,ei(cid:54)=e} e∩ ei. For example, d hH(e1) = 3. We follow [21] to define its degree by d(v) = (cid:80){e∈Ev∈e} w(e). the incidence matrix, d(v) =(cid:80){e∈E} w(e)h(v, e). When all w(e) = 1, d(v) counts the number of hyperedges which include this node: d(v) =(cid:80){e∈Ev∈e} 1, or d(v) = (cid:80){e∈E} h(v, e). Let Dv denote the diagonal matrices containing the node degrees. 4. We then define the hyperdegree of a node as d H(v) :=(cid:80){v∈ee∈E} δ(e), which 3, d hH(e3) = 2, and d hH(e4) = 0 in Figure 1. For a node v ∈ V : If one uses counts the number of nodes connected to a particular node across all hyperedges. For example, d H(v2) = 5, d H(v3) = 6, d H(v5) = 3 in Figure 1. version will be estimating the strength between the connected node pairs. Its weighted Next, let W denote the diagonal matrix containing the weights w(e) of hyper- edges. Correspondingly, the adjacency matrix A of hypergraph G is defined as: 4 (a) Toy example of a hypergraph (b) Simple graph (c) Incidence matrix H Left: Figure 1: Hypergraph and its corresponding simple graph and incidence ma- an hyperedge set E = {e1, e2, e3, e4} and a node set V = trix. {v1, v2, v3, v4, v5, v6, v7}. Middle: the corresponding simple graph. Right: the in- cidence matrix H of the hypergraph on the left, with the entry (vi, ej) being set to 1 if vi is in ej, and 0 otherwise. A = HWHT − Dv, (1) where HT is the transpose of H. 2.2.2 Graphcut of the Hypergraph One can group the nodes into subsets using graph partitioning methods, i.e., graph- cut. The intuition is to find a partition of the graph such that the edges within a adjacency matrix A(X, Y ) :=(cid:80) subset have high weights (strong intra-class connections), and the edges between dif- ferent subsets have low weights (weak inter-class connections). Let S ∈ V denote a subset of nodes and Sc denote the complement of S. Follow the notations in [25], the i∈X,j∈Y aij. For a given number M of subnets, the Mincut approach [26] implements the graphcut by generating a partition S1, . . . , SM which minimizes cut(S1, . . . , SM ) := 1 2 A(Si, Sc i ). (2) M(cid:88) i=1 To solve the problem of separating individual nodes as a subset in Mincut, Ra- tioCut [27] and Normalized cuts (Ncuts) [28] have been proposed to encode the information of the size of a subset. RatioCut(S1, . . . , SM ) := M(cid:88) i=1 1 2 A(Si, Sc i ) Si = M(cid:88) i=1 cut(Si, Sc i ) Si , where S measures the number of nodes in S. Ncut(S1, . . . , SM ) := 1 2 M(cid:88) i=1 A(Si, Sc i ) vol(Si) = 5 M(cid:88) i=1 cut(Si, Sc i ) vol(Si) , (3) (4) attached to the nodes as vol(S) := (cid:80) where vol(S) measures the volume of S by summing over the weights of all edges v∈S ds(v), and node strength ds(v) is the weighted version of node degree d(v). Ncuts has been widely used in image segmentation and brain study commu- nity, since it utilizes the weight information. In the following, we show that Ncuts approach can be generalized from simple graphs to hypergraphs, which has been proven in [21]. For a hypergraph G = (V ; E; w), a cut is a partition of V into two parts S and Sc. A hypergraph e is cut when it is incident with the nodes in S and Sc at the same time. The hyperedge boundary of S is defined as ∂S := {e ∈ Ee∩S (cid:54)= ∅, e ∈ Ee∩Sc (cid:54)= ∅}, which is a hyperedge set consisting of the hyperedges which are cut [21]. The nodes in S, vol(S) := (cid:80) definition of the volume in a hypergraph vol(S) is the sum of the degrees of the v∈S d(v). Each hyperedge is essentially a fully connected subgraph, then the edges in a subgraph is called subedges, being assigned with the same weight w(e)/δ(e). When a hyperdege e is cut, there are e∩ Se∩ Sc subedges are cut. Hence, the volume of ∂S is defined by vol(∂S) := e ∩ Se ∩ Sc δ(e) , w(e) (5) (cid:88) e∈∂S which is the sum of weights over the subedges being cut. By this definition, we have vol(∂S) = vol(∂Sc). Similar to the simple graphs, Normalized hypergraph cut is to keep the high intra-class connection and low inter-class connection with a partition S1, . . . , SM by minimizing the cut as below: M(cid:88) i=1 argmin ∅(cid:54)=S1,...,SM⊂V vol(∂Si) vol(Si) . (6) 2.3 Task Activation Detection - Node Definition in the Hy- pergraph In order to construct the multiple layers in the hypergraph, we apply the activation detection technique on the task data to define the nodes that are contained in different hyperedges. The standard way of activation detection is to use a General Linear Model (GLM) where statistics, such as t-values, reflect the degree of the similarity between the stimulus and voxel time courses. The estimated statistics produce an activation statistics map (t-map), followed by a thresholding of the map to identify the activated voxels [29]. Due to the pronounced noise in the Functional Magnetic Resonance Imaging (fMRI) data, activation detection at the individual level could be inaccurate [30]. In order to derive more reliable task- induced activation, we have chosen a group activation detection over the individual 6 based approach. First, to compute the intra-subject activation patters, a standard GLM is applied as below [29]: Yi = Xiβi + Ei, (7) where Yi is a t×N matrix of the task-induced fMRI time courses of N brain regions from subject i, βi is a d× N activation matrix to be estimated, Ei is a t× N residual confounds] is a t × d matrix. Xtask is the task regressors matrix, and Xi = [XtaskXi and Xi confounds is the confound regressors. Next, we combine the activation results across subjects to assemble a group activation map, which is used to define nodes for each layer of the hypergraph. Specifically, we apply a max-t permutation test [31] on βi aggregated from all the subjects, which implicitly accounts for multiple comparisons and control over false detections [32]. Group activation is declared at a p-value threshold of 0.05. 2.4 Strength Informed Weighted Multi-task Hypergraph In the beginning of section 2, we argued that multi-task information can be pre- sented as a hypergraph, with the hyperedges being different tasks, and the nodes in each hyperedge being the brain regions activated in a certain task. In the tradi- tional definition of hypergraph, nodes are connected to each other binarily, i.e., the edge weights between a node pair are 1 if they are connected, or 0 otherwise. We here propose a strength informed weighted hypergraph model by incorporating the strength information from the connections between nodes. We further determine the hyperedge weight w(e) using the graphical measures defined in subsubsection 2.2.1. 2.4.1 Pairwise Nodal Connection Strength Estimation In order to estimate the strength of the connections between two nodes, we use the Pearson's correlations between time courses from pairs of brain regions. We denote the resting state connectivity matrix as Crest. To produce the task-induced connectivity matrix Ctask, we use the task-induced time course information. We follow the strategy in [8] to remove all inter-block rest periods from all regions' time courses, before computing the pairwise Pearson's correlations across all concatenated block/event duration time courses within a task. To keep the consistency when combining information from the nodes across different layers, we keep all the Ctask having the same dimension of N × N as the Crest, then set the rows and columns of non-activated nodes to zero. 7 2.4.2 Proposed Strength Informed Weighted Hypergraph We present a modified hypergraph cut criteria formulation based on Equation 5 to incorporate pairwise nodal connection strength information from C as below in Equation 8. The symbol . indicates the usage of strength information. (cid:80) (cid:88) e∈∂S vol(∂S) := w(e) i∈{e∩S},j∈{e∩Sc} Ce ij δ(e) , (8) where vol(∂S) is a strength informed version of vol(∂S) in Equation 5, Ce is the connectivity matrix derived from the task corresponding to the layer e, and w(e) is the modified weight item in the hypergraph. We propose here to incorporate strength information from the connectivity matrix and utilize the four hypergraph measures defined in subsubsection 2.2.1 to determine w(e), whose nature is the importance of the hyperedge in the hypergraph. Based on the definition of the four hypergraph measures, we exploit their corresponding biological meanings to set vol(∂S) and w(e) as below: 1. The degree of a hyperedge δ(e) counts the number of brain regions that are activated in a task. To avoid the bias of the hyperedge size, vol(∂S) should be normalized by δ(e). 2. The hyperdegree of a hyperedge is defined as the number of hyperedges that are connected to it. Higher value indicates that more frequently activated patterns in the brain activities exist in this hyperedge. Thus, w(e) should be proportional to d hH(e), i.e., w(e) ∝ d hH(e). 3. The degree of a node counts the number of hyperedges that contain this node, and the biological equivalence is the number of different tasks in which one node is activated. A node with a higher degree is similar to the definition of the connector hubs residing within different subnetworks. Hence, w(e) should be proportional to some statistics derived from d(v) of the nodes in a hyperedge e. We denote the statistics computation method as stat here and it can be widely used statistics such as average value (mean), median value (median) and maximum value (max). Thus, w(e) ∝ stat(d(v)). 4. The hyperdegree of a node reflects the number of all other nodes that are connected to it across all layers, which equals the number of connections from other co-activated nodes to it across multiple tasks. The biological meaning of a node with a high value coincides with the definition of hubs. Hence, w(e) should be proportional to some statistics derived from d H(v) of the nodes in a hyperedge e, i.e., w(e) ∝ stat(d H(v)). Here, in order to incorporate strength information, we apply the weighted version of d H(v), the strength of the node d Hs(v) as defined 8 in Equation 9, i.e., w(e) ∝ stat(d Hs(v)). (cid:88) (cid:88) {v∈ee∈E} u∈e d Hs(v) := Ce uv, (9) where Ce is the task-induced connectivity matrix for the eth task. In order to utilize strength information and hypergraph measures, we propose the w(e) formulation as below: w(e) := w1 · d hH(e) + w2 · stat(d(v)) + w3 · stat(d Hs(v)), (10) where w1, w2, w3 are free parameters to control the contributions of each measure to the hyperedge. 2.5 Multisource Integration of Rest and Task fMRI Given the close correspondence between task and rest connectivity architecture and subnetworks, we further extend the multi-task hypergraph model to integrate Resting State Functional Connectivity based on MRI (rs-fcMRI) information. To do that, we use Crest for the pairwise nodal connection strength computation in Equation 9 as below: d Hs(v) := Crest uv , (11) (cid:88) (cid:88) {v∈ee∈E} u∈e Furthermore, we explicitly combine the two sources of task and rest data for sub- network extraction. We firstly fuse the multiple layers of the multi-task hypergraph into one single layer, and secondly combine it with a resting state connectivity layer. Given that the hypergraph cut criterion (Equation 5) is to evaluate the aggregated sum of the cuts across all the pairwise subedges (nodal connections) in the hyper- gragh, we propose to aggregate the strength information between node pairs across all the layers. To do that, we transform the multiple pairwise nodal connections across task layers (Equation 8) into one single nodal connection as below: T(cid:88) k=1 ¯Ctask ij = 1 T w(ek) δek Cek ij , (12) where the subscript k = 1, . . . , T is the indicator for tasks, T is the total number of tasks available, and ek is the hyperedge in the kth layer of the hypergraph. Cek is the connectivity matrix derived using the time courses in the task k using the procedure described in subsubsection 2.4.1. We next explicitly combine the two sources by a linear weighted combination between the aggregated multi-task connectivity matrix from above (Equation 12) and the resting state connectivity matrix in Equation 13 as below: Ct-r := γ ¯Ctask + (1 − γ)Crest, (13) 9 where γ a free parameter, which can be optimized by cross-validation, or deter- mined by the number of the tasks available. Our linear model for combining two sources, which are both derived from functional modality, was motivated by the study indicating a largely linear superposition of task-evoked signal and resting state modulations in the brain [7]. We also explore combining the two by applying a multislice community detection approach [33], which extends modularity quality function based on the stability of communities under Laplacian dynamics with a coupling parameter ω to control over interslice correspondence of communities. 3 Results We first investigated the similarity of connectivity between resting state and task- general and task-specific connectivity. To evaluate our proposed approaches, we assessed the graphical metric modularity Q value, the inter-subject reproducibil- ity and examined the biological meaning of subnetwork assignments. We applied subnetwork extraction on (1) resting state FC alone, (2) task-induced FC alone, (3) multi-task hypergraph, (4) multi-task hypergraph integrated with resting state connectivity strength, (5) weighted combination of (4) and resting state FC, (6) com- bination of (4) and resting state FC using multislice community detection method [33]. 3.1 Materials We used the resting state fMRI and task fMRI scans of 77 unrelated healthy subjects from the Human Connectome Project (HCP) dataset [34]. Two sessions of resting state fMRI with 30 minutes for each session, and 7 sessions of task fMRI data were available for multisource integration. The seven tasks are working memory (total time: 10:02), gambling (6:24), motor (7:08), language (7:54), social cognition (6:54), relational processing (5:52) and emotion processing (4:32). Preprocessing already applied to the HCP fMRI data includes gradient distortion correction, motion cor- rection, spatial normalization to Montreal Neurological Institute (MNI) space with nonlinear registration based on a single spline interpolation, and intensity normal- ization [35]. Additionally, we regressed out motion artifacts, mean white matter and cerebrospinal fluid confounds, and principal components of high variance voxels using compCor [36]. Next, we applied a bandpass filter with cutoff frequencies of 0.01 and 0.1 Hz for resting state fMRI data. For task fMRI data, we performed similar temporal processing, except a high-pass filter at 1/128 Hz was used. The data were further demeaned and normalized by the standard deviation. We then used the Harvard-Oxford (HO) atlas [37], which has 112 region of interest (ROI)s, 10 to define the brain region nodes. We chose the well-established HO atlas because it sampled from every major brain system, and consists of the highest number of subjects with both manual and automatic labelling technique compared to other commonly used anatomical atlases. Voxel time courses within ROIs were averaged to generate region time courses. The region time courses were demeaned, normalized by the standard deviation. Group level time courses were generated by concatenat- ing the time courses across subjects. The Pearson's correlation values between the region time courses were taken as estimates of FC matrices. Negative elements in all connectivity matrices were set to zero due to the currently unclear interpretation of negative connectivity [38]. For task activation, we applied the activation detection on the seven tasks available following the steps described in subsection 2.3. We summarize here the annotation of the graphs for six methods being evaluated for subnetwork extraction. (1) Resting state FC matrix Crest is used. (2) The task general FC Ctask was generated by concatenating the time courses across all tasks In (3), we use task-specific FC in Equation 9 before the Pearson's correlation. and Equation 10 for each hyperedge, denoted as Chyper-task. We implement (4) by using resting state FC in Equation 9 and Equation 10 as described in subsection 2.5, denoted as Chyper-t-r. For (5), we first generate ¯Ctask ij by using task-specific FC as Cek ij , and resting state Crest to compute w(ek) based on Equation 9 and Equation 10. We next applied our proposed local thresholding [39] on resting state FC Crest to match with the graph density of ¯Ctask at 0.2765, which lies within the normal range of thresholding before subnetwork extraction between [0.2, 0.3] [3]. We then estimate Ct-r using Equation 13. We set free parameters w1, w2, w3 to one, and the stat to median value based on inner cross-validation. For (6), we generated the ¯Ctask and thresholded Crest as the same way as in (5), then the multisource integration is implemented using a multislice approach [33], denoted as Ct-r-multislice. We set ij the weighting for multisource integration γ or coupling parameter ω from 0.01 to 1 at an interval of 0.01. In order to perform fair comparison, Crest in method (1) and Ctask in method (2) have also been local thresholded at the graph density of 0.2765. Method (1) to (5) used Ncuts and (6) used generalized Louvain as the graph partitioning approach. The number of subnetworks was set to seven given that there are seven tasks available to examine if subnetwork assignments can be related to tasks. We note that setting the number of subnetworks is non-trivial as discussed in the previous section that we leave as future work. All statistical comparisons are based on the Wilcoxon signed rank test with significance declared at an α of 0.05 with Bonferroni correction. 11 3.2 Similarity of FC between Resting state and Task data We observed a similarity at Dice Similarity Coefficient (DSC) = 0.7845 between resting state FC and task general FC, which was generated by concatenating the time courses across all different tasks. For seven specific tasks, the corresponding DSC between task-specific FC and task general FC are 0.8971 for emotion process- ing, 0.8557 gambling, 0.8676 for language, 0.9043 for motor, 0.8594 for relational processing, 0.8307 for social cognition, and 0.8751 for working memory. This high similarities confirms the findings in [8] that a set of small but consistent changes common across tasks suggests the existence of a task-general network architecture distinguishing task states from rest. When resting state FC is compared to task-specific FC, the DSC are 0.7193 for emotion processing, 0.7689 for gambling, 0.7390 for language, 0.7067 for motor, 0.7533 for relational processing, 0.7659 for social cognition and 0.7118 for working memory, respectively. The variation of similarities between task-specific and resting state FC around a relatively high average level further confirms that the brain's functional network architecture during task is configured primarily by an intrinsic network architecture which can be present during rest, and secondarily by changes in evoked task-general (common across tasks) and task-specific network [8]. These findings confirms the close relationship between task and rest, and the support for integrating multitask information into resting state based subnetwork extraction. 3.3 Modularity Q Value Modularity Q value has been used to assess a graph partitioning through reflecting the intra- and inter- subnetwork connection structure of a network [9]. We observe that Q values of group level subnetwork extraction for method (1)-(6) are 0.1401, 0.1282, 0.1624, 0.1711, 0.2290 and 0.1905 when γ and ω were selected at the highest inter-subject reproducibility. At the subject-wise level, the modularity Q values estimated from the subnetwork extraction using method (1)-(6) are 0.1397±0.0142, 0.1234±0.0159, 0.2072 ± 0.0199, 0.2094±0.0189, 0.2183±0.0192, and 0.2089±0.0165 respectively, Figure 2. We show that the modularity estimated from subnetworks extracted based on simply concatenating task time courses is lower than using resting state data. Using hypergraph framework (3) Chyper-task and (4) Chyper-t-r achieves statistically higher modularity values than using either resting state data or simple concatenation of task data. Moreover, incorporating resting state information into the hypergraph framework (5) Ct-r can increase modularity compared to hypergraph method. Mul- tislice integration (6) Ct-r-slice results in a lower modularity than (5) the linear model; 12 Figure 2: Subject-wise level modularity Q values using Method (1)-(6). For method (5) and (6), parameter γ and ω were selected at the highest inter-subject repro- ducibility. however, it still outperforms all the other uni-source methods. Overall, incorporat- ing resting state information explicitly using a weighted combination strategy, i.e., method (5) gives a statistically higher modularity than all contrasted methods at p < 10−4 based on Wilcoxon signed rank test. We note that the Q values derived here are around 0.2, when the number of the subnetworks was set to seven, i.e., the number of tasks. It is relatively low due to the inherent resolution limit of Q, i.e., Q decreases when the number of subnetworks increases. We explored this direc- tion by achieving the similar level of Q values around 0.3-0.4 when the number of subnetworks decreases to 4 as in [13]. 3.4 Inter-subject Reproducibility of Subnetwork Extraction We assessed the inter-subject reproducibility by comparing the subnetwork extrac- tion results using subject-wise data against the group level data. The average DSC between subject-wise and group level subnetworks across 77 subjects based on methods (1)-(6) are 0.6362±0.0828, 0.5704±0.0872, 0.7083±0.1094, 0.7258±0.1201, 0.7561±0.1199, and 0.7406±0.0725, Figure 3. We noticed that the reproducibility using resting state FC Crest is higher than simple concatenation of task time courses data Ctask. It could be that there exist great differences in reaction to stimuli from different subjects, and simple concatenation is hard to discover the higher order relationship between canonical network components. On the other hand, analyz- ing multi-task information using hypergraph (3) Chyper-task achieved much higher stability in subnetwork extraction, and incorporating resting information implicitly within the hypergraph (4) Chyper-t-r, or explicit weighted combination (5) Ct-r can even further enhance reproducibility. We note that the weighted combination out- performs multislice integration (6) Ct-r-slice, which is still better than all the other 13 (1)Crest(2)Ctask(3)Chyper-task(4)Chyper-t-r(5)Ct-r(6)Ct-r-multislice00.050.10.150.20.250.30.35 Figure 3: Subject-wise level inter-subject reproducibility of subnetwork extraction using Method (1)-(6). For method (5) and (6), parameter γ and ω were selected at the highest inter-subject reproducibility. uni-source methods. The reason could be that a simple linear model suffices the fusion of task and rest data. Overall, the inter-subject reproducibility derived by (5) Ct-r is statistically higher than all contrasted methods at p < 10−4 based on Wilcoxon signed rank test. 3.5 Biological Meaning We next examined the biological meaning of the subnetworks extracted from method (1) - (6), where γ was set to 0.5 to report the results when resting state and hy- pergraph based multitask information are equally combined as an example. Seven subnetworks were extracted based on the number of tasks available. Method (1) de- tects most of the traditional resting state subnetworks with several false positive and negative detection. The results of method (2) oftentimes combined some important regions from different subnetworks, which lacks biological justifications. Method (3) and (4) generate similar results and both improve the results of method (2) greatly when bringing task dynamics into the subnetwork extraction. Overall, method (5) detects brain regions, which are more biologically meaningful, by combining the in- trinsic network architecture from resting state data and the task dynamics based on high-order hypergraph. We report our findings in details as the following and the visualization of subnetwork extraction results can be found in Figure 4. Using method (1) based on resting state FC alone, subnetwork 1 and 6 are de- tected as left and right side of a combination of Executive Control Network (ECN) and frontoparietal network, which include superior frontal gyrus, middle frontal gyrus, inferior frontal gyrus, posterior supramarginal gyrus, angular gyrus, frontal orbital cortex, and frontal operculum cortex. Method (1) mistakenly classified left inferior lateral occipital cortex and left anterior supramarginal gyrus into the Left 14 (1)Crest(2)Ctask(3)Chyper-task(4)Chyper-t-r(5)Ct-r(6)Ct-r-multislice0.30.40.50.60.70.80.91 (a) Method (1) Crest (b) Method (2) Ctask (c) Method (3) Chyper-task (d) Method (4) Chyper-t-r (e) Method (5) Ct-r (f) Method (6) Ct-r-slice Figure 4: Visualization of subnetworks extraction using methods (1)-(6). The mass center of each ROI is plotted in the MNI space and colorcoded by the membership of seven subnetworks. 15 Executive Control Network (LECN). Anterior supramarginal gyrus is part of the so- matosensory association cortex, which interprets tactile sensory data and is involved in perception of space and limbs location or language processing, thus it should be included in Default Mode Network (DMN) instead of ECN [40]. On the other hand, our proposed method (5) detects both the left and right sides of most of the anterior portion of ECN and posterior supramarginal gyri for subnetwork 1. Using method (5), the left inferior lateral occipital cortex was not include in ECN, which is more accurate. Besides, method (5) clustered anterior supramarginal gyrus symmetrically into subnetwork 6, which includes both sides of Posterior Cingulate Cortex (PCC), precuneus, and angular gyrus, comprising most of the posterior portion of DMN de- fined in [40]. As for method (2), the simple concatenation of multitask time courses, subnetwork 1 consists of frontal medial cortex and only the left side of frontal orbital cortex, and subnetwork 6 consists of most of the anterior portion of ECN, angular gyrus and only the left posterior supramarginal gyrus, which should be symmetri- cally included in DMN. Besides, there are two other ROIs, left subcallosal cortex and left caudate, included in subnetwork 6, which lacks biological meaning. Subnet- work 1 derived from method (3) and (4) both consist of most of the anterior portion of ECN, except that method (3) has two more one-sided frontal areas, which makes (4) more biological meaningful (with symmetric results). Subnetwork 6 of method (3) and (4) both consist of one isolate area: left anterior parahippocampal gyrus, which further indicates that there is need to incorporate resting state information into the multitask based on hypergraph framework. Subnetwork 2 of method (1) includes both sides of Anterior Cingulate Cortex (ACC), caudate, thalamus, putamen and accumbens. Method (5) includes all the same brain regions as method (1) plus one other region, the insula. This subnetwork should be related to the gambling task and emotional processing, which expect to activate ACC [41, 42], ventral striatum (such as thalamus [42] and accumbens [43]), and insula [44]. Usually insula is part of the salience network and has been found to play key roles in emotional processing [45]. However, using method (1), the insula was clustered into subnetwork 5 (mostly motor system). Method (2) included right ACC and both sides of PCC, precuneous, left side of supracalcarine cortex, and accumbens inside subnetwork 2, which seems like a mixture of part of DMN, one- sided region from motor system, and one region from gambling system. As for method (3) and (4), they both extracted similar regions for subnetwork 2 as using method (5), except that they missed thalamus and falsely included left frontal medial cortex. Subnetwork 3 derived from method (1) includes superior lateral occipital cortex, frontal medial cortex, left subcallosal cortex, PCC, precuneous, parahippocampal gyrus, temporal fusiform cortex, brain stem, hippocampus and amygdala. This as- 16 signment does not make too much sense by clustering regions from visual, auditory, emotion circuit and frontal system together. Meanwhile, the results using method (5) consists mostly of emotion circuit and social processing, which includes brain stem [46], hippocampus and parahippocampal gyrus [47], amygdala [48], and sub- callosal cortex [49]. Method (5) also detected regions related to auditory functions such as temporal pole, which is reasonable since the negative emotion was induced by listening to stories. Subnetwork 3 detected by method (2) includes right anterior parahippocampal gyrus, temporal fusiform cortex and brain stem, which still lacks important brain regions in the emotion circuit. Method (3) detects more biologically meaningful regions than (2), such as hippocampus and amygdala. Using method (4) can even detect more related regions than method (3), such as frontal orbital cortex [50]. Method (1) and (5) detected almost the same brain regions for subnetwork 4, which is the visual system, except that method (5) detected one more region of the inferior lateral occipital cortex, making the results more symmetric. This subnet- work includes inferior lateral occipital cortex, intracalcarine cortex, cuneal cortex, lingual gyrus, occipital fusiform gyrus, temporal occipital fusiform cortex, occipital pole, and supracalcarine cortex. Method (2) detected most of the visual regions except for cuneal cortex and the right supracalcarine cortex. Method (3) and (4) detected extra regions in right ECN and auditory system besides all the regions found using (5) in the visual system. Subnetwork 5 derived from method (1) comprises of the motor system, including precentral gyrus, postcentral gyrus, only the right side of anterior supramarginal gyrus, juxtapositional lobule cortex; and the frontoparietal network including left central opercular cortex, superior parietal lobule, and parietal operculum cortex. Method (5) generated similar results as method (1), only that the results are more symmetric, which include both sides of anterior supramarginal gyrus (part of so- matosensory association cortex); and more accurate in terms of frontoparietal net- work, which includes frontal operculum cortex instead of central opercular cortex. Both method (3) and (4) generated similar regions for subnetwork 5 as well, which includes motor system and frontoparietal network, except that they both included brain stem into this subnetwork. However, method (2) mis-classified insula, puta- men and thalamus into the motor and frontal parietal networks. We note that the motor system and frontoparietal network are clustered together, it could be that the working memory tasks recruited both the motor system and frontoparietal network. As for the subnetwork 7, both method (1) and (5) detected brain regions cor- responding to language task and related auditory regions, such as anterior superior temporal gyrus, planum temporale, planum polare, and Heschls gyrus (includes H1 and H2) [51]. Different from method (1), method (5) included central opercular 17 cortex, which can be explained by how fronto-opercular is related to language [52]. Method (2) detected some false positive brain regions in the language system such as parahippocampal gyrus, hippocampus and amygdala. Method (3) and (4) correctly clustered all the brain regions into the language network as method (5). Method (6) generated similar results compared to method (5), only a couple regions in subnetworks 2 and 5 were switched, a couple regions in subnetwork 6 and 7 were switched, and a couple regions in 1 and 6 were switched. Overall, The subnetwork results derived by method (5) Ct-r have more biological meaning than contrasted methods. 4 Discussion 4.1 Hypergraph encodes higher order nodal relationship Subnetwork results derived from methods based on hypergraph achieved higher mod- ularity, higher inter-subject reproducibility, and more reasonable biological meaning than traditional connectivity analysis of pairwise correlation between nodes. These results indicate that hypergraph, which is a natural presentation of multitask acti- vation, can be explored to study higher order relations among the network nodes. The proposed strength informed version of automatic weight setting of the hyper- edge incorporates connectivity information to reveal more accurate higher order relationship among nodes rather than just using binary information. 4.2 Multisource Integration Improves Subnetwork Extrac- tion We have proved that multisource integration of task and rest information can im- prove subnetwork extraction compared to using a single source in terms of graphical metrics, inter-subject reproducibility, along with biologically meaningful subnetwork assignments. We note that the implicit integration of rest information into multi- task hypergraph achieved less improvements as the explicit integration based on the linear combination. The reason could be that the limited number of tasks available restricts the comprehensive representation of the brain using the hypergraph. Thus, by integrating rest data to compensate possible missing information resulted in over- all better outcomes. Another observation is that the linear combination outperforms the multislice community detection, which still performs better than uni-source ap- proaches. Our assumption is that rest and task FC are both derived from a single functional modality, which complements each other by revealing the two sides of FC, i.e., the resting intrinsic side and the activated evoked side. Thus, a simple 18 linear weighted combination would suffice this situation, which outperforms other alternative combination approach in practice. 4.3 Limitations and Future Directions There are several limitations in our present work. First, our study investigated only seven available tasks with high quality data and decent amount of data per task. This sample of seven tasks is not enough. A possible solution is to have access to both task and rest data from previous task studies or co-activation studies, which covers much wider variety of tasks. At the same time, with much more information from a greater amount of task data, we can devise a reliable automatic manner to determine the integration weighting parameter γ. The underlying rationale is that with more tasks available, we can rely more on the hypergraph based multitask source, hence the higher γ. Secondly, we set the number of the subnetworks to be seven, which corresponds to the number of tasks available. The reason is simply to see if we can associate the subnetwork results to different tasks and gain insights from the findings based on task-induced functions. In the future, a finer scale of subnetwork extraction using multi-scale hierarchical approach would improve the interpretation of the findings. 5 Conclusion We proposed a high order relation informed approach based on hypergraph to com- bine the information from multi-task data and resting state data to improve subnet- work extraction. We demonstrated that fusing task activation, task-induced con- nectivity and resting state functional connectivity based on hypergraphs improves subnetwork extraction compared to employing a single source from either rest or task data in terms of subnetwork modularity measure, inter-subject reproducibility, along with more biologically meaningful subnetwork assignments. 6 List of Acronyms ACC Anterior Cingulate Cortex DMN Default Mode Network DSC Dice Similarity Coefficient ECN Executive Control Network FC Functional Connectivity 19 fMRI Functional Magnetic Resonance Imaging GLM General Linear Model HCP Human Connectome Project HO Harvard-Oxford ICA Independent Component Analysis LECN Left Executive Control Network MNI Montreal Neurological Institute Ncuts Normalized cuts PCC Posterior Cingulate Cortex ROI region of interest rs-fcMRI Resting State Functional Connectivity based on MRI SNR signal-to-noise ratio References [1] Park, B., Kim, D.S., Park, H.J.: Graph independent component analysis reveals repertoires of intrinsic network components in the human brain. PloS one 9(1) (2014) e82873 [2] Girvan, M., Newman, M.E.: Community structure in social and biological networks. Proceedings of the National Academy of Sciences 99(12) (2002) 7821 -- 7826 [3] Van Den Heuvel, M., Mandl, R., Pol, H.H.: Normalized cut group clustering of resting-state fmri data. PloS one 3(4) (2008) e2001 [4] Nicolini, C., Bifone, A.: Modular structure of brain functional networks: break- ing the resolution limit by surprise. Scientific reports 6 (2016) [5] Murphy, K., Birn, R.M., Bandettini, P.A.: Resting-state fmri confounds and cleanup. Neuroimage 80 (2013) 349 -- 359 [6] Smith, S.M., Fox, P.T., Miller, K.L., Glahn, D.C., Fox, P.M., Mackay, C.E., Filippini, N., Watkins, K.E., Toro, R., Laird, A.R., et al.: Correspondence of the brain's functional architecture during activation and rest. Proceedings of the National Academy of Sciences 106(31) (2009) 13040 -- 13045 20 [7] Fox, M.D., Snyder, A.Z., Vincent, J.L., Raichle, M.E.: Intrinsic fluctuations within cortical systems account for intertrial variability in human behavior. Neuron 56(1) (2007) 171 -- 184 [8] Cole, M.W., Bassett, D.S., Power, J.D., Braver, T.S., Petersen, S.E.: Intrinsic and task-evoked network architectures of the human brain. Neuron 83(1) (2014) 238 -- 251 [9] Sporns, O., Betzel, R.F.: Modular brain networks. Annual review of psychology 67 (2016) 613 -- 640 [10] Tavor, I., Jones, O.P., Mars, R., Smith, S., Behrens, T., Jbabdi, S.: Task-free mri predicts individual differences in brain activity during task performance. Science 352(6282) (2016) 216 -- 220 [11] Chan, M.Y., Alhazmi, F.H., Park, D.C., Savalia, N.K., Wig, G.S.: Resting- state network topology differentiates task signals across the adult life span. Journal of Neuroscience 37(10) (2017) 2734 -- 2745 [12] Mennes, M., Kelly, C., Zuo, X.N., Di Martino, A., Biswal, B.B., Castellanos, F.X., Milham, M.P.: Inter-individual differences in resting-state functional connectivity predict task-induced bold activity. Neuroimage 50(4) (2010) 1690 -- 1701 [13] Crossley, N.A., Mechelli, A., V´ertes, P.E., Winton-Brown, T.T., Patel, A.X., Ginestet, C.E., McGuire, P., Bullmore, E.T.: Cognitive relevance of the com- munity structure of the human brain functional coactivation network. Proceed- ings of the National Academy of Sciences 110(28) (2013) 11583 -- 11588 [14] Bertolero, M.A., Yeo, B.T., D'Esposito, M.: The modular and integrative func- tional architecture of the human brain. Proceedings of the National Academy of Sciences 112(49) (2015) E6798 -- E6807 [15] Bassett, D.S., Sporns, O.: Network neuroscience. Nature neuroscience 20(3) (2017) 353 [16] Laird, A.R., Fox, P.M., Eickhoff, S.B., Turner, J.A., Ray, K.L., McKay, D.R., Glahn, D.C., Beckmann, C.F., Smith, S.M., Fox, P.T.: Behavioral interpre- tations of intrinsic connectivity networks. Journal of cognitive neuroscience 23(12) (2011) 4022 -- 4037 [17] Mesulam, M.M.: From sensation to cognition. Brain: a journal of neurology 121(6) (1998) 1013 -- 1052 21 [18] De Domenico, M.: Multilayer modeling and analysis of human brain networks. Giga Science 6(5) (2017) 1 -- 8 [19] Muldoon, S.F., Bassett, D.S.: Network and multilayer network approaches to understanding human brain dynamics. Philosophy of Science 83(5) (2016) 710 -- 720 [20] De Domenico, M., Sasai, S., Arenas, A.: Mapping multiplex hubs in human functional brain networks. Frontiers in neuroscience 10 (2016) [21] Zhou, D., Huang, J., Scholkopf, B.: Learning with hypergraphs: Clustering, classification, and embedding. In: Advances in neural information processing systems. (2007) 1601 -- 1608 [22] Dotko, P., Hess, K., Levi, R., Nolte, M., Reimann, M., Scolamiero, M., Turner, K., Muller, E., Markram, H.: Topological analysis of the connectome of digital reconstructions of neural microcircuits. arXiv preprint arXiv:1601.01580 (2016) [23] Zu, C., Gao, Y., Munsell, B., Kim, M., Peng, Z., Zhu, Y., Gao, W., Zhang, D., Shen, D., Wu, G.: Identifying high order brain connectome biomarkers via learning on hypergraph. In: International Workshop on Machine Learning in Medical Imaging, Springer (2016) 1 -- 9 [24] Munsell, B.C., Wu, G., Gao, Y., Desisto, N., Styner, M.: Identifying rela- tionships in functional and structural connectome data using a hypergraph learning method. In: International Conference on Medical Image Computing and Computer-Assisted Intervention, Springer (2016) 9 -- 17 [25] Von Luxburg, U.: A tutorial on spectral clustering. Statistics and computing 17(4) (2007) 395 -- 416 [26] Stoer, M., Wagner, F.: A simple min-cut algorithm. Journal of the ACM (JACM) 44(4) (1997) 585 -- 591 [27] Hagen, L., Kahng, A.B.: New spectral methods for ratio cut partitioning and clustering. IEEE transactions on computer-aided design of integrated circuits and systems 11(9) (1992) 1074 -- 1085 [28] Shi, J., Malik, J.: Normalized cuts and image segmentation. IEEE Transactions on pattern analysis and machine intelligence 22(8) (2000) 888 -- 905 [29] Friston, K.J., Holmes, A.P., Worsley, K.J., Poline, J.P., Frith, C.D., Frack- owiak, R.S.: Statistical parametric maps in functional imaging: a general linear approach. Human brain mapping 2(4) (1994) 189 -- 210 22 [30] Ng, B., Hamarneh, G., Abugharbieh, R.: Modeling brain activation in fmri using group mrf. IEEE transactions on medical imaging 31(5) (2012) 1113 -- 1123 [31] Nichols, T., Hayasaka, S.: Controlling the familywise error rate in functional neuroimaging: a comparative review. Statistical methods in medical research 12(5) (2003) 419 -- 446 [32] Yoldemir, B.: Multimodal fusion for assessing functional segregation and inte- gration in the human brain. PhD thesis, University of British Columbia (2016) [33] Mucha, P.J., Richardson, T., Macon, K., Porter, M.A., Onnela, J.P.: Commu- nity structure in time-dependent, multiscale, and multiplex networks. science 328(5980) (2010) 876 -- 878 [34] Van Essen, D.C., Smith, S.M., Barch, D.M., Behrens, T.E., Yacoub, E., Ugur- bil, K., Consortium, W.M.H., et al.: The wu-minn human connectome project: an overview. Neuroimage 80 (2013) 62 -- 79 [35] Glasser, M.F., Sotiropoulos, S.N., Wilson, J.A., Coalson, T.S., Fischl, B., An- dersson, J.L., Xu, J., Jbabdi, S., Webster, M., Polimeni, J.R., et al.: The minimal preprocessing pipelines for the human connectome project. Neuroim- age 80 (2013) 105 -- 124 [36] Behzadi, Y., Restom, K., Liau, J., Liu, T.T.: A component based noise correc- tion method (compcor) for bold and perfusion based fmri. Neuroimage 37(1) (2007) 90 -- 101 [37] Desikan, R.S., S´egonne, F., Fischl, B., Quinn, B.T., Dickerson, B.C., Blacker, D., Buckner, R.L., Dale, A.M., Maguire, R.P., Hyman, B.T., et al.: An auto- mated labeling system for subdividing the human cerebral cortex on mri scans into gyral based regions of interest. Neuroimage 31(3) (2006) 968 -- 980 [38] Skudlarski, P., Jagannathan, K., Calhoun, V.D., Hampson, M., Skudlarska, B.A., Pearlson, G.: Measuring brain connectivity: diffusion tensor imaging validates resting state temporal correlations. Neuroimage 43(3) (2008) 554 -- 561 [39] Wang, C., Ng, B., Abugharbieh, R.: Modularity reinforcement for improving brain subnetwork extraction. In: International Conference on Medical Image Computing and Computer-Assisted Intervention, Springer (2016) 132 -- 139 23 [40] Heinonen, J., Numminen, J., Hlushchuk, Y., Antell, H., Taatila, V., Suomala, J.: Default mode and executive networks areas: Association with the serial order in divergent thinking. PloS one 11(9) (2016) e0162234 [41] Charpentier, C.J., Martino, B.D., Sim, A.L., Sharot, T., Roiser, J.P.: Emotion- induced loss aversion and striatal-amygdala coupling in low-anxious individuals. Social cognitive and affective neuroscience 11(4) (2015) 569 -- 579 [42] Koehler, S., Ovadia-Caro, S., van der Meer, E., Villringer, A., Heinz, A., Romanczuk-Seiferth, N., Margulies, D.S.: Increased functional connectivity between prefrontal cortex and reward system in pathological gambling. PLoS One 8(12) (2013) e84565 [43] Limbrick-Oldfield, E.H., Mick, I., Cocks, R., McGonigle, J., Sharman, S., Gold- stone, A.P., Stokes, P., Waldman, A., Erritzoe, D., Bowden-Jones, H., et al.: Neural substrates of cue reactivity and craving in gambling disorder. Transla- tional psychiatry 7(1) (2017) e992 [44] Leong, J.K., Pestilli, F., Wu, C.C., Samanez-Larkin, G.R., Knutson, B.: White- matter tract connecting anterior insula to nucleus accumbens correlates with reduced preference for positively skewed gambles. Neuron 89(1) (2016) 63 -- 69 [45] Cauda, F., D'agata, F., Sacco, K., Duca, S., Geminiani, G., Vercelli, A.: Func- tional connectivity of the insula in the resting brain. Neuroimage 55(1) (2011) 8 -- 23 [46] Venkatraman, A., Edlow, B.L., Immordino-Yang, M.H.: The brainstem in emotion: A review. Frontiers in neuroanatomy 11 (2017) [47] Ohmura, Y., Izumi, T., Yamaguchi, T., Tsutsui-Kimura, I., Yoshida, T., Yosh- ioka, M.: The serotonergic projection from the median raphe nucleus to the ventral hippocampus is involved in the retrieval of fear memory through the corticotropin-releasing factor type 2 receptor. Neuropsychopharmacology 35(6) (2010) 1271 [48] Zald, D.H.: The human amygdala and the emotional evaluation of sensory stimuli. Brain Research Reviews 41(1) (2003) 88 -- 123 [49] Laxton, A.W., Neimat, J.S., Davis, K.D., Womelsdorf, T., Hutchison, W.D., Dostrovsky, J.O., Hamani, C., Mayberg, H.S., Lozano, A.M.: Neuronal cod- ing of implicit emotion categories in the subcallosal cortex in patients with depression. Biological psychiatry 74(10) (2013) 714 -- 719 24 [50] Levens, S.M., Devinsky, O., Phelps, E.A.: Role of the left amygdala and right orbital frontal cortex in emotional interference resolution facilitation in working memory. Neuropsychologia 49(12) (2011) 3201 -- 3212 [51] Noesselt, T., Shah, N.J., Jancke, L.: Top-down and bottom-up modulation of language related areas -- an fmri study. BMC neuroscience 4(1) (2003) 13 [52] Meyer, M., Alter, K., Friederici, A.: Functional mr imaging exposes differential brain responses to syntax and prosody during auditory sentence comprehension. Journal of Neurolinguistics 16(4) (2003) 277 -- 300 25
1503.07552
1
1503
2015-03-25T20:54:56
Making a swim central pattern generator out of latent parabolic bursters
[ "q-bio.NC", "nlin.PS" ]
We study the rhythmogenesis of oscillatory patterns emerging in network motifs composed of inhibitory coupled tonic spiking neurons represented by the Plant model of R15 nerve cells. Such motifs are argued to be used as building blocks for a larger central pattern generator network controlling swim locomotion of sea slug Melibe leonina.
q-bio.NC
q-bio
September 25, 2018 10:41 parabolicbursters International Journal of Bifurcation and Chaos © World Scientific Publishing Company 5 1 0 2 r a M 5 2 ] . C N o i b - q [ 1 v 2 5 5 7 0 . 3 0 5 1 : v i X r a MAKING A SWIM CENTER PATTERN GENERATOR OUT OF LATENT PARABOLIC BURSTERS Department of Mathematics and Statistics, Georgia State University, Atlanta 30303, USA. DENIZ ALAC¸ AM Email: [email protected]. ANDREY SHILNIKOV Neuroscience Institute and Department of Mathematics and Statistics, Georgia State University, Atlanta Institute for Information Technology, Mathematics and Mechanics, Lobachevsky State University of Nizhni Novgorod, Nizhni Novgorod, 603950, Russia. 30303, USA. Email: [email protected]. Received (to be inserted by publisher) We study the rhythmogenesis of oscillatory patterns emerging in network motifs composed of inhibitory coupled tonic spiking neurons represented by the Plant model of R15 nerve cells. Such motifs are argued to be used as building blocks for a larger central pattern generator network controlling swim locomotion of sea slug Melibe leonina. Keywords: Plant model, parabolic bursting, half-center oscillation, central pattern generator, swim locomotion, see slug 1. Introduction A plethora of vital rhythmic motor behaviors, such as heartbeat, respiratory functions and locomotion are produced and governed by neural networks called central pattern generators (CPGs) [Selverston, 1985; Bal et al., 1988; Marder & Calabrese, 1996; Frost & Katz, 1996; Kristan et al., 2005; Katz & Hooper, 2007]. A CPG is a microcircuit of interneurons whose mutually synergetic interactions autonomously generate an array of multi-phase bursting rhythms underlying motor behaviors. There is a growing consensus in the community of neurophysiologists and computational researchers that some basic structural and func- tional elements must be shared by CPGs in invertebrate and vertebrate animals. As such, we should first understand these elements, find the universal principles, and develop efficient mathematical and computa- tional tools for plausible and phenomenological models of CPG networks. Pairing experimental studies and modeling studies has proven to be key to unlocking insights into operational and dynamical principles of CPGs [Gillner & Wallen, 1985; Kopell & Ermentrout, 2004; Matsuoka, 1987; Kopell, 1988; Canavier et al., 1994; Skinner et al., 1994a; Dror et al., 1999; Prinz et al., 2003a]. Although various circuits and models of specific CPGs have been developed, it still remains unclear what makes the CPG dynamics so robust and flexible [Best et al., 2005; Belykh & Shilnikov, 2008; Sherwood et al., 2010; Koch et al., 2011; Calabrese et al., 2011; Marder, 2012]. It is also unclear what mechanisms a multi-functional motor system can use to generate polyrhythmic outcomes to govern several behaviors [Kristan, 2008; Briggman & Kristan, 2008; Wojcik et al., 2014]. Our goal is to gain insight into the fundamental and universal rules governing pattern formation in complex networks of neurons. To achieve this goal, we should identify the rules underlying 1 September 25, 2018 10:41 parabolicbursters 2 Ala¸cam and Shilnikov (a) (b) Fig. 1. (a) Melibe leonina lateral swim style. (b) Network bursting in swim interneurons (Si) of the Melibe swim CPG halts when Si3R is hyperpolarized, thus its counterpart Si3L begins tonic spiking; the photographs and in-vitro recording provided courtesy of A. Sakurai [Sakurai et al., 2014] the emergence of cooperative rhythms in simple CPG networks. Recently, a great deal of computational studies have been focused on a range of 3-cell motifs of bursting neurons coupled by chemical (inhibitory and excitatory) and electrical synapses to disclose the role of coupling in generating sets of coexisting rhythmic outcomes, see [Shilnikov et al., 2008; Wojcik et al., 2011, 2014; Schwabedal et al., 2014, 2015; Collens et al., 2015] and references therein. These network structures reflect the known physiological details of various CPG networks in real animals. Next, we would like to explore dynamics and stability of some identified CPG circuits constituted by 4-cells [Jalil et al., 2013]. Examples of such sub-networks can be found in the cerebral crustacean stomatogastric ganglion (STG) [Selverston, 1985; Prinz et al., 2003b, 2004; Marder, 2012], as well as in the swim CPGs of the sea slugs -- Melibe leonina (depicted during swimming in Fig. 1(a)) and Dendronotus iris [Newcomb et al., 2012; Sakurai et al., 2011; Sakurai & Katz, 2011]. Our greater goal is to create dynamical foundations for the onset, morphogenesis and structural robustness of rhythmic activity patterns produced by swim CPGs in these animals. A pilot mathematical model of the Melibe swim CPG will be discussed in this paper. The circuitry shown in Fig. 2(a) depicts only some core elements identified in the biological CPG; its detailed diagram can be found in [Sakurai et al., 2014]. (a) (b) (a) A core circuitry of the biological Melibe swim CPG with inhibitory (●), excitatory (▼) and electrical ((cid:131)~(cid:131)~) synapses Fig. 2. [Sakurai et al., 2014]. (b) In-vitro voltage activity recordings from identified swim interneurons, Si2L and Si3L/R, of the Melibe swim CPG with the characteristic 3 4 -phase lag between the HCO2 and HCO3; intracellular recording provided courtesy of A. Sakurai [Sakurai et al., 2014]. Being inspired by experimental studies of voltage activity recorded from the swim CPGs of the sea slugs Melibe leonina and Dendronotus iris, we would like to develop an assembly line for CPG constructors September 25, 2018 10:41 parabolicbursters CPGs out of latent parabolic bursters 3 Fig. 3. CPG. Recording provided courtesy of A. Sakurai and time series analysis by A. Kelley. (a) Parabolic distribution of spike frequency within bursts produced by networked interneurons in the Melibe swim made of coupled biophysically plausible models. Our first simplifying assumption is that CPGs are made of universal building blocks -- half center oscillations (HCOs) [Hill et al., 2003]. Loosely speaking, a HCO is treated as a pair of interneurons interacting with each other through reciprocally inhibitory synapses and exhibiting anti-phase bursting. The interneurons of a HCO can be endogenous bursters, tonic spiking or quiescent ones, which exhibit alternating bursting only when they inhibit each other. Theoretical stud- ies [Wang & Rinzel, 1985] have indicated that formation of an anti-phase bursting rhythm is always based on slow subsystem dynamics. There are three basic mechanisms to generate alternating bursting in the HCO: release, escape, and post-inhibitory rebound (PIR). The first mechanism is typical for endogenously bursting neurons [Jalil et al., 2010, 2012]. The other two mechanisms underlie network bursting in HCOs comprised of neurons, which are hyperpolarized quiescent in isolation [Perkel & Mulloney, 1974; Skinner et al., 1994b; Angstadt et al., 2005; Kopell & Ermentrout, 2002]. Our second assumption is that the swim CPG interneurons are intrinsic tonic spikers that become network bursters only when externally driven or coupled by inhibitory synapses, as recent experimental studies suggest [Sakurai et al., 2014]. The third assumption is that network bursting in the Melibe swim CPG is parabolic, i.e. the spike frequency within a burst increases at the middle, and decreases at the ends, as one can observe from Fig. 3. This observation indicates the type of neuronal models to be employed to describe network cores. Our model of choice for parabolic bursting is the Plant model [Plant & Kim, 1975, 1976; Plant, 1981]. The Plant model has been developed to accurately describe the voltage dynamics of the R15 neuron in a mollusk Aplysia Californica, which has turned out to be an endogenous burster [Levitan & Levitan, 1988]. Most dynamical properties of the R15 neuron have been modeled and studied in detail [Canavier et al., 1991; Bertran, 1993; Butera et al., 1995; Butera, 1998; Sieling & Butera, 2011; Ji et al., 2013]. 2. Methods: the Plant model of parabolic bursting The conductance based Plant model [Plant, 1981] for the R15 neuron [Sieling & Butera, 2011] located in the abdominal ganglion of a slug Aplysia Californica is given by the following set of ordinary differential equations derived within the framework of the Hodgkin-Huxley formalism to describe the dynamics of the fast inward sodium [Na], outward potassium [K], slow TTX-resistant calcium [Ca] and an outward calcium sensitive potassium [KCa] currents: Cm V =−IN a− ICa− IKCa− Ileak− Iext− Isyn. (1) The last three currents are the generic ohmic leak Ileak, external constant Iext and synaptic Isyn currents flowing from a pre-synaptic neuron. The full details of the representation of the currents employed in the model are given in the Appendix below. September 25, 2018 10:41 parabolicbursters 4 Ala¸cam and Shilnikov (a) (b) (c) There are two bifurcation parameters in the individual model. The first one is the constant external introduced in the slowest equation: Ca= ρ(Kcx(VCa− V + ∆)− Ca) (2) Fig. 4. (a) Endogenous bursting in the Plant model as alternations of tonic spiking activity and quiescent periods. (b) Single burst featuring a characteristic spike frequency increase in the middle of each burst. (c) Parabolic shape of the frequency distribution of spikes within a burst is a feature of this kind of bursting. The parameters are ρ= 0.00015ms−1, Kc= 0.00425ms−1 and τx= 9400ms. current, Iext, which is set Iext = 0. Following [Shilnikov, 2012], the other bifurcation parameter, ∆, is deviation from a mean value of the reversal potential VCa = 140mV evaluated experimentally for the At ∆= 0, the neuron is an endogenous burster, see Fig. 4. According to [Rinzel & Lee, 1987], this calcium current in the R15 cells. As such, this makes ∆ a bifurcation parameter. Secondly its variations are not supposed to alter the topology of the slow motion manifolds in the 5D phase space, which are called tonic spiking and quiescent in the mathematical neuroscience context, as they are made of, respectively, round periodic orbits and equilibrium states [of the slow subsystem] of the model (Fig. 5). describing the concentration of the intracellular calcium in the Plant model. By construction, ∆ is a type of bursting is termed parabolic. The reason for this term is that the spike frequency within bursts is maximized in the middle of bursts and minimized at the beginning and the end (see Fig. 4c). The parabolic structure of a burst is due to the calcium-activated potassium current. Its magnitude is determined by the intracellular calcium concentration. As the intracellular calcium concentration increases, the calcium Fig. 5. Bursting (green) orbit recursively switching between two slow -- motion critical manifolds: tonic spiking, Mlc, with a characteristic fold and originating through a sub-critical Andronov-Hopf (AH) bifurcation from a depolarized equilibrium state, and quiescent, Meq (orange curve), projected onto the (h, V ) and slow Ca variables of the of the Plant model; a plane represents the synaptic threshold, Θsyn= 0mV . September 25, 2018 10:41 parabolicbursters CPGs out of latent parabolic bursters 5 Fig. 6. Responses of the bursting neuron (∆= 0mV ) on the synaptic drive Isyn= gsyn(V − Vrev). (a) Excitatory synaptic drive with gsyn= 0.002nS and Vrev= 40mV applied at t= 80sec switches the neuron from bursting to tonic spiking activity. (b) The inhibitory drive with gsyn= 0.005nS and Vsyn=−80mV halts bursting and makes the neuron hyperpolarized quiescent. dependent potassium current gets activated, which causes an increase of the inward potassium current. As the membrane potential increases over a threshold value, the intracellular calcium concentration decreases, as well as the inward potassium current (see Eq. (20) in the Appendix). The parabolic distribution of spikes within bursts is shown in Fig. 4. The instant frequency value is calculated by the reciprocal of each inter-spike interval. Panels b and c of Fig. 4 clearly disclose the parabolic inter-spike structure of bursts. It was shown in [Rinzel & Lee, 1987] that the mechanism underlying a transition between quiescent and tonic spiking of bursting in the Plant model is due to a homoclinic bifurcation of a saddle-node equilibrium state [Shilnikov, 1963; Afraimovich et al., 2014]. This bifurcation occurs in the fast 3D(V, h, n)-subspace of the model and is modulated by the 2D slow dynamics in the(Ca, x)-variables, which are determined by slow oscillations of the intracellular calcium concentration [Plant & Kim, 1975, 1976]. The unfolding of this codimension-one bifurcation includes an onset of a stable equilibrium, which is associated with a hyperpolarized phase of bursting, and on the other end, an emergent stable periodic orbit that is associated with tonic spiking phase of bursting. The period of this stable orbit decreases, as it moves further away from the saddle-node equilibrium mediated by decreasing calcium concentration. The period of the tonic spiking orbit grows with no upper bound as it approaches the homoclinic loop of the saddle-node [Shilnikov et al., 1998,2001]. Variations of ∆ change the duty cycle of bursting, which is a ratio of the active tonic spiking phase of bursting to its period. Decreasing ∆ reduces the inactive, quiescent phase of bursting, i.e. increases its duty cycle. Zero duty cycle is associated with the homoclinic saddle-node bifurcation that makes the neuron hyperpolarized quiescent. This corresponds to an emergence of stable equilibrium state for all dynamical variables of the model (1). In other words, decreasing ∆ makes the active phase longer, so that below a threshold ∆=−32mV the neuron switches to tonic spiking activity. Tonic spiking activity is associated with the emergence of a stable periodic orbit in the fast(V, h, n)-subspace, while the(Ca, x)-variables of the slow due to relaxation of periodic oscillations in the 2D (Ca, x)-subspace, which slowly modulates fast tonic spiking oscillations in the(V, h, n) variables. The relaxation limit cycles emerge from one and collapse into the other equilibrium state in the(Ca, x)-plane through Andronov-Hopf bifurcations, which can be sub- subspace converge to a stable equilibrium state. As such, bursting occurs in the Plant and similar models or super-critical. At the transitions between bursting and tonic spiking, and bursting and hyperpolarized quiescence, the neuron can produce chaotic dynamics, which are basically due to the membrane potential oscillatory perturbations of plain canards at the folds of the relaxation cycle. (a) (b) September 25, 2018 10:41 parabolicbursters 6 Ala¸cam and Shilnikov Fig. 7. Tonic spiking neuron 1 at ∆=−34mV near the bifurcation transition between tonic spiking and bursting is forced to syn= 0.001nS, from the pre-synaptic neuron 2 at t= 60sec. Halting the inhibitory drive restores tonic spiking activity in the targeted neuron (not shown). become a network burster with an application of an inhibitory drive with ginh 3. Endogenous and network bursting. Inhibitory and excitatory drives A half-center oscillator is a network of two neurons coupled by reciprocally inhibitory synapses that robustly produces bursting in alternation, or anti-phase bursting. Such a network can be multistable, i.e. produce other bursting rhythms as well, such as synchronous bursting [Jalil et al., 2010] and rhythmic outcomes with slightly shifted phase lags between the endogenously bursting neurons [Jalil et al., 2012]. In this study, the synaptic current Isyn is modeled through the fast threshold modulation (FTM) approach [Kopell & Somers, 1993]. The synapses are assumed to be fast and non-delayed, which is true for the swim CPG in both sea slugs under consideration. The synaptic current is given by Isyn= gsyn(Vpost− Esyn) 1+ e−k(Vpre−Θsyn) , 1 (3) where gsyn is the maximal conductance of the current, which is used as a bifurcation parameter of the networked model; Vpost(t) and Vpre(t) are the voltages on the post-synaptic (driven) and pre-synaptic (driving) neurons; Esyn is the synaptic reversal potential. To make Isyn excitatory, we set Esyn= 40mV , while in the inhibitory case we set Esyn=−80mV . In Eq. (3), the second term is a Boltzmann coupling function that quickly, (k= 100), turns the synaptic current on and off as soon the voltage, Vpre, of the (driving) pre-synaptic cell(s) raises above and falls below the synaptic threshold, here Θsyn= 0mV (Fig. 5). To model the constant synaptic drive onto the post-synaptic neuron, we assume that Vpre > Θsyn. This allows us to calibrate the state of the post-synaptic neuron, and to determine the drive threshold that separates the qualitatively distinct states of the individual and networked neurons. This statement is illustrated in Fig. 6 by simulating responses of the endogenous parabolic burster to network perturbation. Figure 6(a) shows, with a properly adjusted excitatory drive, that the endogenous burster switches into tonic spiking activity. On the other hand, bursting in the networked neuron can be halted when it receives a sufficient inhibitory drive from the pre-synaptic neuron of the network (Figure 6(b)). Eliminating either drive makes the post-synaptic neuron return to its natural state, i.e. these experiments de-facto prove that the neuron is mono-stable for the given parameter values. A HCO, in the canonical Brown definition [Graham-Brown, 1911], is a pair of neurons bursting in anti-phase when they are networked by inhibitory synapses. In isolation, such neurons are not endogenous bursters but tonic spikers instead, or remain quiescent [Marder & Calabrese, 1996]. There are multiple mechanisms underlying such anti-phase bursting, or, more accurately, anti-phase oscillations in HCOs and CPGs made of relaxation oscillators [Kopell & Ermentrout, 2002; Daun et al., 2009]. The list includes the well studied mechanisms of post-inhibitory rebound and escape for quiescent neurons [Perkel & Mulloney, 1974; Wang & Rinzel, 1985; Skinner et al., 1994b; Destexhe et al., 1994; Matveev et al., 2007], as well as less-known mechanisms of HCOs constituted by intrinsically spiking neurons. Such networks utilizing the Plant models are discussed below. into the tonic spiking mode. This is done by setting the bifurcation parameter, ∆=−34mV , see Fig. 7. To construct such a HCO with relatively weak inhibitory coupling, the Plant model must be first set September 25, 2018 10:41 parabolicbursters CPGs out of latent parabolic bursters 7 Fig. 8. Anti-phase network bursting produced by a HCO of two Plant neurons as soon as the inhibition is turned on. Blocking the inhibition restores tonic spiking activity in both neurons, and vice versa. Here, the the network parameters are ginh syn = 0.008nS and Esyn =−80mV , and the parameters of the individual neurons are the following: ∆=−60mV, ρ= 0.0003ms−1, Kc= 0.0085ms−1, τx= 235ms and x∞(V)= 1~(1+ e−0.15(V+50)). drive of gsyn= 0.001nS from the post-synaptic neuron 2 at t= 60sec. The inhibitory drive is sufficient to Next, we consider a unidirectional network where the tonic spiking neuron 1 starts receiving, an inhibitory shift the post-inhibitory neuron over the bifurcation transition back into bursting activity. The minimal inhibitory drive must be increased proportionally to make the targeted neuron a network burster whenever it stays further away from the bifurcation transition between tonic spiking and bursting in isolation. 4. Forming a half-center oscillator In this section, we discuss the dynamics of half-center oscillators made of two tonically spiking Plant neurons reciprocally coupled by inhibitory synapses. As before, we describe such synapses within the framework of the fast threshold modulation (FTM) paradigm using Eq. (3) to match the shape and magnitude of inhibitory postsynaptic potentials (IPSPs) in the post-synaptic neurons. IPSPs are the indicators of the type and the strength of synapses in the network. We perform simulations in a fashion that is analogous to the dynamic clamp technique used in neu- rophysiological experiments. The approach involves the dynamic block, restoration and modulation of synaptic connections during simulation. These modeling perturbations should closely resemble the exper- imental techniques of drug-induced synaptic blockade, modulation, wash-out, etc. Restoring the chemical synapses during a simulation makes the HCO regain network bursting activity with specific phase charac- teristics. Depending on the coupling strength as well as the way the tonically spiking neurons are clamped, the network bursting may change phase-locked states, i.e. be potentially multi-stable. Experimental ob- servations also suggest specific constraints on the range of coupling strengths of the reciprocal inhibition, such that the networks stably and generically achieve the desired phase-locking. are initiated in tonic spiking mode. After turning on the reciprocally inhibitory synapses gsyn= 0.008nS, Figure 8 demonstrates the stages of anti-phase bursting formation in the HCO. The uncoupled neurons the HCO quickly transitions to the regime of robust anti-phase bursting. Turning off the synapses restores the native tonic spiking activity in both neurons. Turning on the reciprocal synapses makes the HCO regain the network bursting. Note that the length of transients from tonic spiking to network bursting depends on the strength of the synaptic coupling for the fixed parameters of the individual Plant neurons. By comparing the magnitude of IPSPs in the voltage traces represented in Figs. 8 and 9, one can conclude that the coupling in the later case is weaker. This is why, the onset of network bursting in the HCO is less pronounced. Our modeling studies agree well with experimental recordings from the identified interneurons in the Melibe swim CPG which suggests that the observed bursting is due to synergetic interactions of interneurons of the network [Sakurai et al., 2014]. One can see from Fig. 1(b) that network bursting in the biological HCO formed by two Si3 interneurons of the Melibe swim CPG is seized as soon as the right one, Si3R, September 25, 2018 10:41 parabolicbursters 8 Ala¸cam and Shilnikov Fig. 9. Onset of emergent network anti-phase bursting in the HCO with reciprocally inhibitory. (Esyn=−80mV ) synapses syn= 0.0073nS. at ginh receives a negative current pulse that makes it hyperpolarized quiescent, while its left bursting counterpart, Si3L, turns into tonic spiking activity instead. Moreover, one can deduct from the wiring diagram of the CPG depicted in Fig. 2(a) and the analysis of voltage traces represented in Fig. 1(b) that the interneuron Si2L becomes a tonic-spiker as soon as the pre-synaptic interneuron Si3R stops inhibiting it (compare with Fig. 8.) This further supports the assertion that the swim CPG is made of intrinsically tonic spiking interneurons. To test the robustness of network anti-phase bursting to perturbation and to calibrate the necessary influx of reciprocal inhibition generated by the Plant neurons, we consider a HCO with excitatory autapses. The objective here is to determine an equivalent amount of excitatory drive to be projected onto the post- inhibitory network burster to cancel out the inhibitory drive and shift it back to the initial tonic-spiking mode. An autapse is a synapse of a neuron onto itself, where the axon of the neuron ends on its own dendrite. After their discovery [Van der Loos & Glaser, 1972] autapses have been observed in a range of nervous systems. The autapses are arguably to be responsible for tuning of neural networks. This particular con- figuration of the HCO depicted in Fig. 10 is formally motivated by the swim CPG circuitry, see Fig. 2(a). One can see from it that the interneurons of the bottom HCO receive excitatory drives from the top in- terneurons forming the top HCO. We would like to find the threshold over which the neurons no longer form a stably bursting HCO. This would allow us to calibrate and quantify the relative strengths of the mixed synaptic connections in the swim CPG models. Both autapses are introduced to the model using the FTM approach with Eaut= 40mV . In this experiment, In this HCO configuration, each neuron inhibits its counterpart and self-excites through the autapse. Fig. 10. Turning on the excitatory autapses at gexc bursting. aut= 0.016nS in the HCO with ginh syn= 0.0073nS halts pronounced network September 25, 2018 10:41 parabolicbursters CPGs out of latent parabolic bursters 9 Fig. 11. Assembly line of the Melibe swim CPG model out of four intrinsically tonic spiking Plant neurons. First the reciprocal inhibition between Si3R and Si3L is turned on, followed by turning on the reciprocal inhibition between Si2R and Si2L, and next simultaneous turning on unidirectional cross-lateral inhibition from Si3R(L) projected onto Si2L(R), and bi-lateral excitation originating from Si2R(L) down onto Si3R(L). After a short transient, the CPG model exhibits the desired 3/4 phase shift lag between Si2L and Si3L. Compare with voltage traces of the biological CPG in Fig. 2(b). the conductance values for inhibitory synapses are set at ginh generate robust anti-phase bursting as seen in Fig. 9. Next, we add the autapses along with inhibition and gradually increase gexc the network stops exhibiting anti-phase bursting. We note that unlike a permanent excitatory drive from pre-synaptic neurons, an introduction of the excitatory autapse, acting only when the self-driving neuron is above the synaptic threshold, is effectively perturbation equivalent for the calibration purpose. aut proportionally increase the delay. At gexc aut. We found that increasing gexc syn= 0.0073nS. This is sufficient for the HCO to aut= 0.016nS, 5. Assembly line of a Melibe swim CPG In this final section, we put together a pilot model of the Melibe swim CPG according to a circuitry based on identified interneurons and synapses; its wiring diagram is sketched in Fig. 2(a). This network model is made of the two HCOs constituted by tonic spiking Plant neurons. We would like to find out whether this sample CPG model can already produce phase lags similar to those between bursting interneurons in the biological CPG. For the sake of simplicity, we do not include Si4R/L interneurons in the model and we also omit electrical synapses. It is known from experimental studies [Sakurai et al., 2014] that blocking chemical, inhibitory and excitatory synapses between the interneurons may be sufficient to break down the motor pattern by the network. Figure 2(b) points out that the interneurons of either HCO burst in anti-phase and there is the characteristic 3/4 phase lag between the burst initiation in the neurons Si2L and Si3L, as well as between Si2R and Si3L. This phase lag is repeatedly observed in both adult and juvenile animals. As before, we use the Plant neurons initiated in the tonic spiking mode, relatively close to the transition to bursting. Initial conditions of the neurons are randomized. After letting the neurons settled down to tonic spiking activity, the network connections are turned on. As Fig. 11 shows, with the reciprocal inhibition being first turned on, the bottom interneurons Si3L and Si3R become anti-phase network bursters, and so do Si2R and Si2L as soon as the reciprocal inhibition between them is turned them on, too. At this stage, the CPG model is formed by two uncoupled HCOs. A few seconds later, they become coupled by September 25, 2018 10:41 parabolicbursters 10 Ala¸cam and Shilnikov simultaneous turning on the unidirectional cross-lateral inhibition from Si3R(L) projected onto Si2L(R), and bi-lateral excitation from Si2R(L) down onto Si3R(L). One can see from this figure that all four interneurons of the CPG model exhibit network bursting with the desired phase lags. These are 0.5 (half period) between the interneurons of each HCO, and 3/4 (a fraction of the network period) between the HCOs, or between the corresponding reference interneurons Si2L and Si3L. We note that such a phase shift was reported in a similar Melibe swim CPG constituted by endogenous bursters; that model also incorporated electrical synapses [Jalil et al., 2013]. There is a great room for improvement of CPG network models to include other identified interneurons and to incorporate additional electrical synapses to find out whether additions of new elements can stabilize or desynchronize the desired bursting pattern as it was done using the Poincar´e return maps for endogenous bursters [Wojcik et al., 2014]. Of our special interest are various problems concerning structural stability of the network, and its robustness (Lyapunov stability) for bursting outcomes subjected to perturbations by pulses of the external current, as well as reductions to return maps between burst initiations in constituent neurons. These questions are beyond the scopes of the given examination and will be addressed in full detail in our forthcoming publications soon. The question about a possible linking of the characteristic 3 4 phase lag and the Melibe leonina lateral swim style is the paramount one among them. 6. Summary We have discussed a basic procedure for building network bursting CPGs made of intrinsically tonic spiking neurons. As a model for such networks, we have employed the biophysically plausible Plant model that was originally proposed to describe endogenous bursting R15-cells in the Aplysia mollusk. Such bursting was intracellularly recorded, and identified as parabolic, from the known interneurons in the swim CPGs of two sea slugs: Melibe leonina and Dendronotus iris. There is experimental evidence that bursting in these swim CPGs is due to synergetic interactions of all constituent neurons that are intrinsic tonic-spikers in isolation. To model the Melibe swim CPG, we have first examined dynamical and structural properties of the Plant model and its responses to perturbations. These perturbations include inhibitory and excitatory inputs from pre-synaptic neurons in the network. We have identified the transition boundary beyond which the bursting Plant model becomes a tonic-spiker and shifted it slightly over the threshold using an introduced bifurcation parameter. We have shown that the perturbed/calibrated Plant neuron, exhibiting intrinsically tonic spiking activity, becomes a network burster when it receives an inhibitory drive from a pre-synaptic neuron. By combining two such neurons, we have created a genuine half-center oscillator robustly producing anti-phase bursting dynamics. We have also considered a HCO configuration with two excitatory autapses to assess the robustness of anti-phase bursting with respect to excitatory perturbations. Finally, we have employed all necessary components to assemble a truncated model of the Melibe swim CPG with the characteristic 3/4-phase lags between the bursting onsets in the four constituent interneurons. In future studies, we plan to examine the dynamics of the CPG models with all synaptic connections, including electrical, as well as incorporating additional identified interneurons. We will also explore their structural stability, robustness and potential multi-stability of their bursting outcomes with various phase lags. An additional goal is to find out whether the motor pattern with the 3/4-phase lags will persist in networks with interneurons represented by other mathematical models including phenomenologically reduced ones. Potentially, these findings shall provide a systematic basis for comprehension of plausible biophysical mechanisms for the origination and regulation of rhythmic patterns generated by various CPGs. Our goal is to extend and generalize the dynamical principles disclosed in the considered networks for other neural systems besides locomotion, such as olfactory cellular networks. 7. Acknowledgments A.S. also acknowledges the support from GSU Brain and Behaviors pilot grant, RFFI 11-01-00001, RSF grant 14-41-00044 at the Lobachevsky University of Nizhny Novgorod and the grant in the agreement of August 27, 2013 N 02.B.49.21.0003 between the Ministry of education and science of the Russian Federation and Lobachevsky State University of Nizhni Novgorod (Sections 2-4), as well as NSF BIO-DMS grant XXXXX. We thank A. Sakurai and P.S. Katz for sharing experimental data and suggestions, and the September 25, 2018 10:41 parabolicbursters acting and associated members of the NEURDS (Neuro Dynamical Systems) lab at GSU: A. Kelley, K. Pusuluri, S. Pusuluri, T.Xing, M. Fen, F. Fen, J. Collens, D. Knapper, A. Noriega, P. Ozluk, B. Chung and E. Latash for helpful discussions, and R. Clewley for his guidance on the PyDSTool package [Clewley et al., 2006] used in simulations. CPGs out of latent parabolic bursters 11 September 25, 2018 10:41 parabolicbursters 12 Ala¸cam and Shilnikov 8. Appendix: the conductance based Plant model Cm V =−IN a− IK− ICa− IKCa− Ileak− Isyn, The model in this study is adopted from [Plant, 1981]. The dynamics of the membrane potential, V , is governed by the following equation: where Cm= 1µF~cm2 is the membrane capacitance, IN a is the N a+ current, IK is the K+ current, ICa is the Ca+2 current, IKCa is the Ca2+ activated K+ current, Ileak is the leak current, Isyn is the synaptic where the reversal potential VN a = 30mV and the maximum N a+ conductance value gN a = 4nS. The current. The fast inward sodium current is given by (4) (5) instantaneous activation variable is defined as IN a= gN am3∞(V)h(V − VN a), αm(V) m∞(V)= αm(V)+ βm(V) , 50− Vs βm(V)= 4 exp((25− Vs)~18), , while the dynamics of inactivation variable h is given by The fast potassium current is given by the equation where the reversal potential is VK=−75mV and the maximum K+ conductance value is gK= 0.3nS.The (12) dynamics of inactivation gating variable is described by (6) (7) (8) (9) (10) (11) (13) (14) (15) (16) where where with where where with αm(V)= 0.1 exp((50− Vs)~10)− 1 h= h∞(V)− h τh(V) and τh(V)= αh(V)= 0.07 exp((25− Vs)~20) and βh(V)= αh(V) αh(V)+ βh(V) h∞(V)= , 12.5 αh(V)+ βh(V) , exp((55− Vs)~10)+ 1 1 , 105 mV. Vs= 127V + 8265 IK= gKn4(V − VK), n= n∞(V)− n τn(V) and τn(V)= and βn(V)= 0.125 exp((45− Vs)~80). αh(V)+ βh(V) , 12.5 , ICa= gCax(V − VCa), αh(V) n∞(V)= αh(V)+ βh(V) 55− Vs αn(V)= 0.01 exp((55− Vs)~10)− 1 The TTX-resistant calcium current is given by September 25, 2018 10:41 parabolicbursters REFERENCES 13 where governed by , 1 x∞(V)= dynamics of the slow activation variable is described by where the reversal potential is VCa= 140mV and the maximum Ca2+ conductance is gCa= 0.03nS. The x= x∞(V)− x τx(V) and τx(V)= 9400ms. exp(−0.3(V + 40))+ 1 The outward Ca2+ activated K+ current is given by [Ca]i (V − VK), IKCa= gKCa 0.5+[Ca]i where the reversal potential is VCa = 140mV . The dynamics of intracellular calcium concentration is Ca= ρ[Kcx(VCa− V)−[Ca]i] , where the reversal potential is VCa = 140mV , and the constant values are ρ= 0.00015mV−1 and Kc = 0.00425mV−1. The leak current is given by Ileak= gL(V − VL), where the reversal potential VL=−40mV and the maximum conductance value gL= 0.0003nS. The synaptic Isyn= gsyn(Vpost− Erev) 1+ e−k(Vpre−Θsyn) with the synaptic reversal potential Vpost=−80mV for inhibitory synapses and Vpost= 40mV for excitatory synapses and the synaptic threshold Θsyn= 0mV , and k= 100. (20) (21) current is defined as (17) (18) (19) (22) References Afraimovich, V., Gonchenko, S., Lerman, L., Shilnikov, A. & Turaev, D. [2014] "Scientific heritage of L.P. Shilnikov. Part 1." Regular and Chaotic Dynamics 19, 435 -- 460. Angstadt, J., Grassmann, J., Theriault, K. & Levasseur, S. [2005] "Mechanisms of postinhibitory rebound and its modulation by serotonin in excitatory swim motor neurons of the medicinal leech," Journal of Comparative Physiology A-Neuroethology, Sensory, Neural and Behavioral Physiology 191, 715 -- 732, doi:10.1007/s00359-005-0628-6. Bal, T., Nagy, F. & Moulins, M. [1988] "The pyloric central pattern generator in crustacea: a set of conditional neural oscillators," Journal of Comparative Physiology A 163, 715 -- 727. Belykh, I. & Shilnikov, A. [2008] "When weak inhibition synchronizes strongly desynchronizing networks of bursting neurons." Phys Rev Lett 101, 078102. Bertran, R. [1993] "A computational study of the effects of serotonin on a molluscan burster neuron," Biol. Cybern. 69, 257 -- 267. Best, J., Borisyuk, A., Rubin, J., Terman, D. & M., W. [2005] "The dynamic range of bursting in a model respiratory pacemaker network," SIAM J. Appl. Dyn. Syst. 4, 1107 -- 1139. Briggman, K. L. & Kristan, W. B. [2008] "Multifunctional pattern-generating circuits." Annu Rev Neurosci 31, 271 -- 294, doi:10.1146/annurev.neuro.31.060407.125552, URL http://dx.doi.org/10.1146/annurev. neuro.31.060407.125552. Butera, R. [1998] "Multirhythmic bursting," Chaos 8, 274 -- 284. Butera, R., Clark, J. W. J., Canavier, C. C., Baxter, D. A. & Byrne, J. H. [1995] "Analysis of the effects of modulatory agents on a modeled bursting neuron: Dynamic interactions between voltage and calcium dependent systems," J Comput. Neuroscience 2, 19 -- 44. Calabrese, RL., Norris, BJ., Wenning, A. & Wright, TM. [2011] "Coping with variability in small neuronal networks," Integrative and Comparative Biology 51, 845 -- 855, doi:10.1093/icb/icr074. September 25, 2018 10:41 parabolicbursters 14 REFERENCES Canavier, C., Clark, J. & Byrne, J. [1991] "Simulation of the bursting activity of neuron R15 in Aplysia: Role of ionic currents, calcium balance, and modulatory transmitters" J Neurophysiology 66, 2107 -- 2124. Canavier, C., Baxter, DA., Clark, JW. & Byrne, JH. [1994] "Multiple modes of activity in a model neuron suggest a novel mechanism for the effects of neuromodulator" J Neurophysiology 72, 872 -- 882. Clewley, R., Sherwood, W., LaMar, M. & Guckenheimer, J. [2006] "Pydstool: an integrated simulation, modeling, and analysis package for dynamical systems." Tech. rep., http://pydstool.sourceforge.net. Collens, J., Kelley, A., Alacam, D., Xing, T., D., K., Schwabedal, J. & Shilnikov, A. [2015] "Intrinsic mechanisms for pattern generation in three-node networks," submitted . Daun, S., Rubin, J. E. & Rybak, I. A. [2009] "Control of oscillation periods and phase durations in half-center central pattern generators: a comparative mechanistic analysis," Journal of Com- putational Neuroscience 27, 3 -- 36, doi:10.1007/s10827-008-0124-4, URL http://dx.doi.org/10.1007/ s10827-008-0124-4. Destexhe, A., Contreras, D., Sejnowski, T. & Steriade, M. [1994] "A Model of Spindle Rhythmicity in the Isolated Thalamic Reticular Nucleus," Journal of Neurophysiology 72, 803 -- 818. Dror, R. O., Canavier, C. C., Butera, R. J., Clark, J. W. & Byrne, J. H. [1999] "A mathematical criterion based on phase response curves for stability in a ring of coupled oscillators." Biol Cybern 80, 11 -- 23, doi:10.1007/s004220050501, URL http://dx.doi.org/10.1007/s004220050501. Frost, W. N. & Katz, P. S. [1996] "Single neuron control over a complex motor program," Proc.Nat.Acad. Sc. 93, 422 -- 426. Gillner, S. & Wallen, P. [1985] "Central pattern generators for locomotion, with special references to vertebrates," Ann. Rev. Neurosci. 8, 233 -- 261. Graham-Brown, T. [1911] "The intrinsic factors in the act of progression in the mammal," Lond B Biol Society 84, 308 -- 319. Hill, A., Van Hooser, S. & Calabrese, R. [2003] "Half-center oscillators underlying rhythmic movements," The Handbook of Brain Theory and Neural Networks, ed. Arbib, M. A. (The MIT Press). Jalil, S., Allen, D., Youker, J. & Shilnikov, A. [2013] "Toward robust phase-locking in melibe swim central pattern generator models," Chaos 23, 046105, doi:http://dx.doi.org/10.1063/1.4825389, URL http: //scitation.aip.org/content/aip/journal/chaos/23/4/10.1063/1.4825389. Jalil, S., Belykh, I. & Shilnikov, A. [2010] "Fast reciprocal inhibition can synchronize bursting neurons," Physical Review E 81, doi:{10.1103/PhysRevE.81.045201}. Physical Review E 85, doi:{10.1103/PhysRevE.85.036214}. Jalil, S., Belykh, I. & Shilnikov, A. [2012] "Spikes matter for phase-locked bursting in inhibitory neurons," Ji, L., Zhang, J., Lang, X. & Zhang, X. [2013] "Coupling and noise induced spiking-bursting transi- tion in a parabolic bursting model," Chaos: An Interdisciplinary Journal of Nonlinear Science 23, 013141, doi:http://dx.doi.org/10.1063/1.4795281, URL http://scitation.aip.org/content/aip/journal/ chaos/23/1/10.1063/1.4795281. Katz, P. S. & Hooper, S. [2007] "Invertebrate central pattern generators," Invertebrate Neurobiology, eds. North, G. & R. Greenspan, R. (Cold Spring Harbor Laboratory Press, NY, New York). Koch, H., Garcia, A. J. & Ramirez, J.-M. [2011] "Network reconfiguration and neuronal plasticity in rhythm-generating networks," Integrative and Comparative Biology 51, 856 -- 868, doi:10.1093/icb/ icr099. Kopell, N. [1988] "Toward a theory of modelling central pattern generators," Neural Control of Rhythmic Movements in Vertebrates, eds. Cohen, A., Rossingol, S. & Grillner, S. (Wiley, New York). Kopell, N. & Ermentrout, B. [2004] "Chemical and electrical synapses perform complementary roles in the synchronization of interneuronal networks," Proc. Natl. Acad, Sci. 101, 15482 -- 15487, doi:10.1073/ pnas.0406343101. Kopell, N. & Ermentrout, G. [2002] "Mechanisms of phase-locking and frequency control in pairs of coupled neural oscillators," Handbook on Dynamical Systems 2, 3 -- 54. Kopell, N. & Somers, D. [1993] "Rapid synchronization through fast threshold modulation," Biol. Cybern. 68, 5. Kristan, W. [2008] "Neuronal decision-making circuits." Curr Biol 18, R928 -- R932, doi:10.1016/j.cub.2008. September 25, 2018 10:41 parabolicbursters REFERENCES 15 07.081, URL http://dx.doi.org/10.1016/j.cub.2008.07.081. Kristan, W., Calabrese, R. & Friesen, W. [2005] "Neuronal control of leech behavior," Prog. Neurobiol. 76, 279. Levitan, E. & Levitan, I. B. [1988] "Serotonin acting via cyclic amp enhances both the hyperpolarizing and depolarizing phases of bursting pacemaker activity in the aplysia neuron R15," J Neuroscience 8, 1152 -- 1161. Marder, E. [2012] "Neuromodulation of neuronal circuits: back to the future," Neuron 76, 1. Marder, E. & Calabrese, R. [1996] "Principles of rhythmic motor pattern generation," Physiological Reviews 76, 687 -- 717. Matsuoka, K. [1987] "Mechanisms of frequency and pattern control in the neural rhythms generators," Matveev, V., Bose, A. & Nadim, F. [2007] "Capturing the bursting dynamics of a two-cell inhibitory network using a one-dimensional map," Journal of Computational Neuroscience 23, 169 -- 187, doi: Biol. Cybernetics 1, 1. {10.1007/s10827-007-0026-x}. Newcomb, J., Sakurai, A., Lillvis, J., Gunaratne, C. & Katz, P. S. [2012] "Homology and homoplasy of swimming behaviors and neural circuits in the nudipleura (mollusca, gastropoda, opistho-branchia)," Proc. Natl. Acad. Sci. 109, 10669 -- 76. Perkel, D. & Mulloney, B. [1974] "Mechanism of postinhibitory rebound in molluscan neurons," Science 185, 181 -- 183, doi:10.1126/science.185.4146.181. Plant, R. [1981] "Bifurcation and resonance in a model for bursting nerve cells," J. Math. Biol 11, 15 -- 32. Plant, R. & Kim, M. [1975] "On the mechanism underlying bursting in the aplysia abdominal ganglion r15 cell," Math. Biosci. 26, 357 -- 375. Plant, R. E. & Kim, M. [1976] "Mathematical description of a bursting pacemaker neuron by a modification of the hodgkin-huxley equations," Biophys J 16, 227 -- 244. Prinz, A., Billimoria, C. & Marder, E. [2003a] "Alternative to hand-tuning conductance-based models: construction and analysis of databases of model neurons," J.Neurophysiol. 90, 3998 -- 4015, URL PM: 12944532. Prinz, A. A., Bucher, D. & Marder, E. [2004] "Similar network activity from disparate circuit parameters," Nature neuroscience 7, 1345 -- 1352. Prinz, A. A., Thirumalai, V. & Marder, E. [2003b] "The functional consequences of changes in the strength and duration of synaptic inputs to oscillatory neurons," The Journal of neuroscience 23, 943 -- 954. Rinzel, J. & Lee, Y. S. [1987] "Dissection of a model for neuronal parabolic bursting," Journal of Mathe- matical Biology 25, 653 -- 675. Sakurai, A., Gunaratne, C. A. & Katz, P. S. [2014] "Two interconnected kernels of reciprocally inhibitory interneurons underlie alternating left-right swim motor pattern generation in the mollusk melibe leonina," Journal of Neurophysiology 112, 317 -- 1328. Sakurai, A. & Katz, P. S. [2011] "Distinct neural circuit architectures produce analogous rhythmic behaviors in related species," Soc. Neurosci. Abstr. 37.918.04. Sakurai, A., Newcomb, J., Lillvis, J. & Katz, P. S. [2011] "Synaptic patterning of left-right alternation in a computational model of the rodent hindlimb central pattern generator," Curr. Biol. 21, 1036. Schwabedal, J., Knapper, D. & Shilnikov, A. [2015] "Qualitative and quantitative stability of polyrhythmic circuits," submitted . Schwabedal, J. T. C., Neiman, A. B. & Shilnikov, A. L. [2014] "Robust design of polyrhythmic neural circuits," Phys. Rev. E 90, 022715, doi:10.1103/PhysRevE.90.022715, URL http://link.aps.org/doi/ 10.1103/PhysRevE.90.022715. Selverston, A. (ed.) [1985] Model Neural Networks and Behavior (Springer, Berlin). Sherwood, W., Harris-Warrick, R. & Guckenheimer, J. [2010] "Synaptic patterning of left-right alternation in a computational model of the rodent hindlimb central pattern generator," J. Comput Neuroscience 30, 323. Shilnikov, A. [2012] "Complete dynamical analysis of a neuron model," Nonlinear Dynamics 68, 305 -- 328, doi:{10.1007/s11071-011-0046-y}. Shilnikov, A., Gordon, R. & Belykh, I. [2008] "Polyrhythmic synchronization in bursting networking mo- September 25, 2018 10:41 parabolicbursters 16 REFERENCES tifs," Chaos 18, doi:{10.1063/1.2959850}. Shilnikov, L. [1963] "Some cases of generation of periodic motion from singular trajectories." Math. USSR Sbornik 61, 443 -- 466. Shilnikov, L., Shilnikov, A., Turaev, D. & Chua, L. [1998,2001] Methods of Qualitative Theory in Nonlinear Dynamics, Parts I and II (World Scientific Publ.), ISBN 9789810240721. Sieling, F. & Butera, R. [2011] "Aplysia R15 neuron," Scholarpedia 6, 4181, URL doi:10.4249/scholarpedia. 4181. Skinner, F., Kopell, N. & Marder, E. [1994a] "Mechanisms for oscillation and frequency control in networks of mutually inhibitory relaxation oscillators," Comput. Neurosci. 1, 69. Skinner, F., Kopell, N. & Marder, E. [1994b] "Mechanisms for oscillation and frequency control in recip- rocally inhibitory model neural networks," Journal of Computational Neuroscience 1, 69 -- 87. Van der Loos, H. & Glaser, E. [1972] "Autapses in neocortex cerebri: synapses between a pyramidal cell's axon and its own dendrites," Brain Research 48. Wang, X.-J. & Rinzel, J. [1985] "Alternating and synchronous rhythms in reciprocally inhibitory model neurons," Ann. Rev. Neurosci. 8, 233 -- 261. Wojcik, J., Clewley, R., Schwabedal, J. & Shilnikov, A. [2014] "Key bifurcations of bursting polyrhythms in 3-cell central pattern generators," PLoS ONE 9. Wojcik, J., Clewley, R. & Shilnikov, A. [2011] "Order parameter for bursting polyrhythms in multifunctional central pattern generators," Phys Rev E 83, 056209 -- 6. Wojcik, J., Schwabedal, J., Clewley, R. & Shilnikov, A. [2014] "Key bifurcation of bursting polyrhythms in 3-cell central pattern generators," PLoS ONE 9, e92918.
1107.2521
3
1107
2013-06-01T14:49:02
Network algorithmics and the emergence of synchronization in cortical models
[ "q-bio.NC" ]
When brain signals are recorded in an electroencephalogram or some similar large-scale record of brain activity, oscillatory patterns are typically observed that are thought to reflect the aggregate electrical activity of the underlying neuronal ensemble. Although it now seems that such patterns participate in feedback loops both temporally with the neurons' spikes and spatially with other brain regions, the mechanisms that might explain the existence of such loops have remained essentially unknown. Here we present a theoretical study of these issues on a cortical model we introduced earlier [Nathan A, Barbosa VC (2010) Network algorithmics and the emergence of the cortical synaptic-weight distribution. Phys Rev E 81: 021916]. We start with the definition of two synchronization measures that aim to capture the synchronization possibilities offered by the model regarding both the overall spiking activity of the neurons and the spiking activity that causes the immediate firing of the postsynaptic neurons. We present computational results on our cortical model, on a model that is random in the Erd\H{o}s-R\'enyi sense, and on a structurally deterministic model. We have found that the algorithmic component underlying our cortical model ultimately provides, through the two synchronization measures, a strong quantitative basis for the emergence of both types of synchronization in all cases. This, in turn, may explain the rise both of temporal feedback loops in the neurons' combined electrical activity and of spatial feedback loops as brain regions that are spatially separated engage in rhythmic behavior.
q-bio.NC
q-bio
1 Network Algorithmics and the Emergence of Synchronization in Cortical Models Andre Nathan, Valmir C. Barbosa∗ Programa de Engenharia de Sistemas e Computa¸cao, COPPE, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil ∗ Corresponding author, e-mail: [email protected] Abstract When brain signals are recorded in an electroencephalogram or some similar large-scale record of brain activity, oscillatory patterns are typically observed that are thought to reflect the aggregate electrical activity of the underlying neuronal ensemble. Although it now seems that such patterns participate in feedback loops both temporally with the neurons' spikes and spatially with other brain regions, the mechanisms that might explain the existence of such loops have remained essentially unknown. Here we present a theoretical study of these issues on a cortical model we introduced earlier [Nathan A, Barbosa VC (2010) Network algorithmics and the emergence of the cortical synaptic-weight distribution. Phys Rev E 81: 021916]. We start with the definition of two synchronization measures that aim to capture the synchronization possibilities offered by the model regarding both the overall spiking activity of the neurons and the spiking activity that causes the immediate firing of the postsynaptic neurons. We present computational results on our cortical model, on a model that is random in the Erdos-R´enyi sense, and on a structurally deterministic model. We have found that the algorithmic component underlying our cortical model ultimately provides, through the two synchronization measures, a strong quantitative basis for the emergence of both types of synchronization in all cases. This, in turn, may explain the rise both of temporal feedback loops in the neurons' combined electrical activity and of spatial feedback loops as brain regions that are spatially separated engage in rhythmic behavior. Introduction Current technology allows the recording of electromagnetic brain activity at several different spatial scales, ranging from invasive recordings that capture the signals from neuronal ensembles comprising thousands to millions of cells to those that are noninvasive (like the electroencephalogram) and capture the signals from large-scale cortical areas [1]. Invariably the recorded signals appear as oscillatory patterns which, depending on the scale at which the recording is performed, occur in different frequency ranges and provide distinct interpretive possibilities. At the smallest spatial scales, for example, the recorded signals occur in the range of a few kilohertz and are thought to reflect the combined spiking (or firing) activity of the group of neurons in question. The signals recorded at the largest scales, on the other hand, occur in the range of a few to 100 hertz and are thought to reflect the so-called local field potentials (LFPs), which in turn are purported to reflect the aggregate effects of all the electrical activity taking place in the corresponding area [2]. LFPs have acquired great prominence recently, owing mainly to the fact that LFPs related to different brain areas appear to combine with one another in such ways as to correlate significantly with the brain's sensory-motor mechanisms and other, higher-level functions as well (such as memory, attention, and others; cf. [3 -- 5] and references therein). The ways in which LFPs combine seem to involve forms of cross- frequency coupling that cover wide temporal and spatial ranges and ultimately promote the integration of activities with different temporal and spatial characteristics [5]. Decoding the various forms of LFP coupling may one day hold the key to understanding how computation and communication take place in the brain. 2 As it happens, though, many aspects of the nature of LFPs have remained elusive. In particular, their precise origin and relation to the underlying firing activity of the neurons seem to depend crucially on the brain region being considered. However, notwithstanding this indefiniteness that still surrounds the specifics of LFP emergence and interaction with other LFPs, some of the crucial points are beginning to be clarified [2, 6]. One of them is the relation with the combined firing activity of groups of neurons. Although LFPs seem to derive largely from the accumulation of potential at the neurons' membranes that eventually leads to neuronal firing, it appears that the firing patterns themselves have a role to play as well. As a consequence, a picture that seems to be emerging is that of a feedback loop in which the larger-scale LFPs both influence and get influenced by the smaller-scale firing patterns. Another point is more related to the spatial characteristics of how LFPs interact with one another: It now seems clear that feedback loops also occur involving the LFPs of distinct brain areas. While a complete resolution of these issues will certainly require considerable further research, a detailed understanding will undoubtedly benefit from a clear picture of firing synchronization at the neuronal level. By synchronization we mean both the convergence of multiple action potentials onto distinct cells within a relatively narrow temporal window, and also the firing due to such potentials within a similar window. As we mentioned above, both phenomena are closely related to the rise of LFPs, and consequently the rise of the higher-level functions that LFPs are thought to support. In our view, accounting for the possibilities the brain affords for the appearance of such synchronized behavior depends both on the anatomical properties of how neurons interconnect and on the individual firing behavior of each neuron. Our aim in this paper is the study of these types of synchronism on a graph- theoretic model of the cortex. Our study is preceded by a few others [7 -- 9], all of which have modeled neuronal behavior as a continuous-time process in which the relevant signals obey certain differential equations and feed some measure of synchronous activity. Invariably these studies have modeled the synchronization of membrane potentials directly, and have for this reason stayed apart from explicitly modeling most of the relevant details that characterize neuronal firings, like the so-called local histories of each neuron [10]. Our ap- proach is to take a different course and focus on the extent to which certain events at the various neurons can be said to be synchronized. These events are, in essence, the arrival of action potentials from other neurons and the eventual creation of new action potentials. These, of course, are precisely the events that make up membrane potentials, but we have found in previous studies of related problems that the additional level of detail pays off by providing unprecedented insight. Examples here are the emergence of the synaptic-weight distribution, in which case we demonstrated that experimentally observed distribu- tions [11] can be reproduced [12], and a more tractable version [13] of the integrated-information theory for the emergence of consciousness [14]. Our model is based on what is generally called network algorithmics. In our case this refers to the combination of a structural component to represent the anatomical properties of the cortex at the neuronal level, and a distributed-algorithmic component to represent the traffic of action potentials among the neurons as they fire. In general, a distributed algorithm is a collection of sequential procedures, each run by an agent in a distributed system with provisions for messages to be sent to other agents for communication [15]. The distributed algorithm of interest here is inherently asynchronous, meaning that the behavior of each agent (each neuron) is purely reactive to the arrival of messages (action potentials), which can in turn undergo unpredictable delays before delivery. It follows that independent runs of the same algorithm from the same initial conditions cannot in general be guaranteed to generate the same sequence of message arrivals at any agent, and therefore not the same behavior. Lately, this inherent source of unpredictability has led asynchronous distributed algorithms to be recognized as a powerful abstraction in the modeling of biological systems [16, 17], since they allow the incorporation into the model of small behavioral variations that might otherwise be simply smoothed out and rendered irrelevant. 3 An asynchronous distributed algorithm never makes any reference to a global-time entity. In our model, therefore, every property emerging from neuronal interactions is the result of strictly local ele- ments. The presence of such strong asynchronism may seem contradictory with the search for signs of synchronism. However, we have found that observing the system from the outside can reveal surprising synchronization possibilities that could never be known within the scope of individual neurons, provided we uncover the causal relationship that binds those local elements on a global scale. Using network algorithmics as the basis of our model makes it substantially more detailed than those that target membrane potentials as the basic analytical units, in the sense that a neuron's behavior can now be described as its reaction to the arrival of a new action potential given its current state. Clearly, though, there would be room for considerably more detail, even down to the molecular level. Our approach, therefore, resides in the middle ground between two extremes and as such is related to what has become known as the artificial-life approach [18, 19], whose "life as it could be" metaphor is close in spirit to our own view. Our approach is also part of a growing collection of efforts that attempt to tackle eminent problems of neuroscience from the perspective of graph-based and other interdisciplinary methods [20 -- 31]. As an enterprise that eventually seeks to be able to handle realistically large problem instances from a graph-theoretic perspective, the present study is moreover in line with the many others that in the last decade have addressed the so-called complex networks [32 -- 34]. Our model is defined on a strongly connected directed graph S, i.e., a directed graph in which it is possible to reach every node from every other node through directed paths only. We assume that S has N nodes, each representing a neuron, so each one can be either excitatory or inhibitory in the proportion of 4 to 1, respectively [35 -- 37]. In S, an edge exists directed from node i to node j to indicate a synapse from i's axon to one of j's dendrites, but no edge can connect two inhibitory nodes [35]. The results we describe in the remainder of the paper for our cortical model are predicated upon S being drawn from the following random-graph model on n ≥ N nodes in such a way that S is the giant strongly connected component, GSCC [38], of the directed graph D that is actually drawn. In D a randomly chosen node, say i, has out-degree k > 0 with probability proportional to k−1.8.1 Following [50, 51], each of these k nodes is precisely another randomly chosen node, say j, with probability proportional to e−dij , where dij is the Euclidean distance between nodes i and j when all n nodes are placed uniformly at random on a radius-1 sphere. We have observed this policy for edge placement to lead to N ≈ 0.9n on average (i.e., we expect the GSCC of D, and hence S, to comprise roughly 90% of the n nodes) [13]. Methods We start with the specification of the asynchronous distributed algorithm to be executed by the nodes of graph S. This algorithm prescribes a common procedure for each node to react to the arrival of a message and to possibly send messages out as a consequence. Such a procedure runs atomically (i.e., without being interrupted), so actually implementing the algorithm requires provisions for the queuing of incoming messages at each node. Each run of the distributed algorithm can be described formally by specifying how the various message arrivals at the different nodes relate to one another. As we do this a precise notion of causality emerges and we use it to define our two synchronization measures. 1Originally the choice of this scale-free distribution [39] was inspired by related work that observed such a power law at a higher level of detail in the cortical architecture [40, 41], but recent measurements regarding the distribution of neuronal (undirected) degrees [42] seem to support the choice we made. Specifically, ignoring edge directions in our model leads to an exponential degree distribution [13], as observed in [42]. We also note that our adoption of a power law to describe the distribution of out-degrees represents a significant departure from earlier modeling attempts [35,43], which used randomness in the Erdos-R´enyi sense [44] directly. We note, moreover, that cortices do exhibit many other scale-free properties [29,45,46], including some that evince their small-world nature [47], the existence of hubs (nodes with very large out-degrees), and others [48, 49]. 4 Asynchronous distributed algorithm All nodes in S share the same rest and threshold potentials, denoted by v0 and vt, respectively, such that v0 < vt. We use vj to represent the potential of node j and wij to represent the synaptic weight of the edge directed from i to j. At all times, these are constrained in such a way that vj ∈ [v0, vt] and wij ∈ [0, 1]. An asynchronous distributed algorithm, henceforth referred to as A, underlies the node- potential and synaptic-weight dynamics of our cortical model. At node j, algorithm A is built around the "firing" operation which consists of sending a message to each out-neighbor of j in S and setting vj to the rest potential v0. Algorithm A is purely reactive, that is, it only acts at any node when the node receives a message. Evidently there must be exceptions to this purely reactive character, since at least one node must send messages out spontaneously to start the algorithm. We call such nodes initiators and let m ≤ N be the number of initiators in a given run of algorithm A. When node j acts as an initiator (this happens at most once in a run for that node), it simply fires with probability 1. When it reacts to the reception of a message, say from node i, node j performs the following steps: 1. If i is excitatory, then set vj to min{vt, vj + wij}. If it is inhibitory, then set vj to max{v0, vj − wij}. 2. Fire with probability (vj − v0)/(vt − v0). 3. If firing did occur during step 2, then set wij to min{1, wij + δ}. If it did not occur but the previous message received by node j from any of its in-neighbors did cause firing to occur, then set wij to (1 − α)wij . The parameters δ and α appearing in step 3 are constrained by δ ≤ α [12]. They are meant to reproduce, although to a limited extent, the spike-timing-dependent plasticity principles [52, 53], which roughly dictate that wij is to increase if firing occurs and decrease otherwise, always taking into account how close in time the relevant firings by nodes i and j are.2 Moreover, increases in wij are to occur by a fixed amount, decreases by proportion [54 -- 56]. Every run of algorithm A is guaranteed to eventually reach a point at which no more message traffic exists. It is then said to have terminated. Modeling causality The problem of handling synchronization during a run of an asynchronous distributed algorithm such as algorithm A is akin to that of handling simultaneity in special relativity. This is so because, in either context, establishing the notion of simultaneity is dependent upon how signals propagate and also on locality considerations [57]. A convenient framework within which to tackle it is then the event-based formalism originally laid down in [58]. An event in the present context is said to have occurred either when an initiator fires (and thereby sends a message to each of its out-neighbors in graph S) or when a node reacts to the reception of a message by executing steps 1 -- 3 (which includes the possibility of firing). In either occasion the event involves the execution of an atomic set of actions by the node in question. A run of algorithm A can then be regarded as a set R of events. An event r ∈ R is described by the 4-tuple r = hi, ti, mi, Mii, where i identifies the node of S at which event r occurred, ti ≥ 1 indicates that r was the tith event to occur at node i since the run began, mi is the message (if any) that triggered the occurrence of the event, and Mi is the set of messages sent by node i if it fired during the event. Many of the events in R are implicitly related to one another. We make such a relation explicit by defining the binary relation B ⊆ R2. Given events r as above and r′ = hj, tj, mj, Mji, we say that the ordered pair (r, r′) ∈ B if and only if: (a) Either i and j are the same node and ti < tj with no intervening event between r and r′; 2In Hebbian terms, learning depends on how the synapses incoming to j compete with one another for the firing of j. Strengthening the synapse from i, in particular, depends on whether the arrival of an action potential from i succeeds in making j fire, either by itself or combined with other recent arrivals that did not cause firing [53]. 5 (b) Or j is an out-neighbor of i and mj ∈ Mi (i.e., the message that triggered r′ was sent in connection with the occurrence of r). We interpret (r, r′) ∈ B as meaning that r happened at node i immediately before r′. If (a) is the case, then this use of "before" seems natural because both ti and tj are relative to the same (local) time basis. Extending this interpretation to case (b), though apparently less natural, is still consistent because there is no reference in this case to either ti or tj. Closing B under transitivity yields another binary relation on R, here denoted by B+, such that B ⊆ B+ ⊆ R2. This relation can be interpreted in such a way that (r, r′) ∈ B+ means that r happened before r′ regardless of how close the specific nodes at which they occurred are to each other in graph S. Through B+, therefore, relation B is the core of a characterization of how the events of run R relate to one another causally. In particular, if r and r′ are such that neither (r, r′) ∈ B+ nor (r′, r) ∈ B+, then the situation is analogous to the space-like separation of events in special relativity (since neither could the occurrence of r influence that of r′ during run R, nor conversely). Relation B is also instrumental in our characterization of synchronization in run R. The first step is to recognize that it naturally gives rise to a directed graph, call it P , whose node set is R (the set of events) and whose edge set is B. This graph is necessarily acyclic (no directed cycles are formed) and allows the definition of event r's depth, denoted by depth(r), as follows. Given a directed path in graph P between two events, let the path's length be defined as the number of edges on the path that correspond to messages. That is, the edges contributing to the path's length are those that fall under case (b) above in the definition of relation B. We define depth(r) as the length of the lengthiest path leading to r in graph P . Intuitively, depth(r) is the size of the longest causal chain of messages leading up to event r during run R. These notions are illustrated in Figure 1. Measures of synchronization We use two measures of synchronization. They are both based on the premise that, if the sequence of events at two neurons (nodes of S) can be aligned with each other so that sufficiently many event pairs of the same depth become matched in the alignment, then there is more synchronism between the two sequences than there would be with fewer event pairs aligned. We do this sequence alignment as described next. Let i and j be the two neurons in question and let their event sequences during j , e2 run R be e1 j , respectively, where Li and Lj are the sequences' sizes. Let µij = max{depth(eLi j )}, i.e., µij is the depth of the last event of the two sequences that is deep- est. We do the alignment of the two sequences by creating two new size-µij sequences, viz. t1, t2, . . . , tµij and u1, u2, . . . , uµij for the first measure, and x1, x2, . . . , xµij and y1, y2, . . . , yµij for the second mea- sure, each with defining characteristics that depend on the particular measure of synchronization under consideration. and e1 i ), depth(eLj j , . . . , eLj i , e2 i , . . . , eLi i The first measure aims to capture the synchronization that may exist in the overall flow of messages and, through it, in the accumulation of potential at the neurons' membranes. As such, it is based on positioning, relative to the t sequence, those of the Li events of neuron i that promoted the accumulation of potential, and likewise the Lj events of neuron j relative to the u sequence. For k = 1, 2, . . . , µij , the t sequence is defined recursively as follows. If k = 1: • tk = 1, if depth(e1 i ) = 1; • tk = 0, otherwise. If k > 1: • tk = k, if depth(eℓ i ) = k for some ℓ ∈ {1, 2, . . . , Li}; • tk = tk−1, otherwise. 6 The u sequence is defined entirely analogously for neuron j. If, for example, the events in the two original sequences have depths 2, 3, 3, 7, 8, 9, 11 and 1, 3, 4, 5, 5, 9 for i and j, respectively, then the t sequence is and the u sequence is t = h0, 2, 3, 3, 3, 3, 7, 8, 9, 9, 11i u = h1, 1, 3, 4, 5, 5, 5, 5, 9, 9, 9i. (1) (2) Thus, the kth position of sequence t or u equals k′ such that 0 < k′ ≤ k if and only if the corresponding neuron received at least one message of depth k′ during the run and, for k′ < k, never since did the neuron receive a message whose depth is in the interval [k′ + 1, k]. It equals 0 otherwise. The first measure of synchronization is denoted by ρ− ij for neurons i and j. It is given by ρ− ij = 1 µij µij Xk=1 min{tk, uk} max{tk, uk} (3) (assuming 0/0 ≡ 1). Clearly, ρ− sequences we get ρ− ij = 1. The example above yields ij ∈ [0, 1] and grows with the similarity of sequences t and u. For identical ρ− ij = 1 11 (cid:18) 0 1 + 1 2 + 3 3 + 3 4 + 3 5 + 3 5 + 5 7 + 5 8 + 9 9 + 9 9 + 9 11(cid:19) ≈ 0.69. (4) The second measure addresses the synchronization possibilities that may exist as the neurons fire. In this case only those of the Li events of neuron i that entailed firing are positioned relative to the x sequence, and likewise the Lj events of neuron j with respect to the y sequence. For k = 1, 2, . . . , µij , the x sequence is such that: • xk = k, if depth(eℓ i ) = k for some ℓ ∈ {1, 2, . . . , Li} such that neuron i fired at the occurrence of eℓ i ; • xk = 0, otherwise. The y sequence is defined analogously for neuron j. Following up on the examples given above, and assuming that neuron i fired at one of its depth-3 events and neuron j fired at all its events except those of depth 5, the x sequence is and the y sequence is x = h0, 0, 3, 0, 0, 0, 0, 0, 0, 0, 0i y = h1, 0, 3, 4, 0, 0, 0, 0, 9, 0, 0i. (5) (6) Thus, the kth position of sequence x or y equals k > 0 if and only if the corresponding neuron fired upon receiving a depth-k message. It equals 0 otherwise. For neurons i and j, the second measure of synchronization is denoted by ρ+ ij and given by ρ+ ij = 1 µij µij Xk=1 min{xk, yk} max{xk, yk} (7) (assuming 0/0 ≡ 1). As in the previous case, ρ+ y. Identical sequences yield ρ+ ij = 1. For the example above we get ij ∈ [0, 1] and grows with the similarity of sequences x and ρ+ ij = 1 11 (cid:18) 0 1 + 0 0 + 3 3 + 0 4 + 0 0 + 0 0 + 0 0 + 0 0 + 0 9 + 0 0 + 0 0(cid:19) ≈ 0.73. (8) ij and ρ+ 7 For a fixed pair i, j of neurons, both ρ− ij seek to characterize the possibility of synchronized behavior during run R. They do this by quantifying the extent to which the events occurring at the two neurons could be said to happen synchronously if a time basis existed common to neurons i and j and time according to this basis elapsed along increasing event depth. Of course, our cortical model per se has no need for such a time basis and is algorithmically correct regardless of any timing assumptions one may make concerning the delay for action-potential propagation down the axons. However, the kind of synchronized behavior we are investigating occurs at a variety of temporal and spatial scales, therefore some time-related assumption is inevitable if we are to extract any meaning out of the multitude of neuronal events. Our choice seems reasonable because, conceptually, it contrasts with the algorithm's inherent asynchronism only minimally: Every single pair of neurons is given its own time basis for the computation of ρ− ij and this is reflected on the notationally explicit dependency of µij on the i, j pair. Furthermore, viewing increasing event depth as time may even have some biological plausibility to it. In fact, it appears that the delay for an action potential to reach the various synapses connecting out of the same axon is independent of how much of the axon actually has to be traversed [59, 60]. In these terms, what our assumption does is to generalize this independence for a group of axons. ij and ρ+ Notwithstanding this common underlying feature of the two synchronization measures, they are also markedly different in how they use event depth to assess the similarity of two event sequences. In the case of ρ+ ij, this is done rather stringently, since only same-depth firing events and the 0/0 ratios contribute to it. In the case of ρ− ij, these continue to be some of the strongest contributors, but now they are joined by any pair of same-depth events (not necessarily firing ones). Moreover, now an event's depth lingers until a greater-depth event occurs and in the meantime continues to influence ρ− ij . Results Our computational results are based on the same methodology we followed in [13], of which we now provide a brief review for the reader's benefit. We use v0 = −15, vt = 0, δ = 0.0002, and α = 0.04 in all runs of algorithm A on S. Each run starts at m = 50 initiators chosen uniformly at random and progresses until termination. A run is implemented as a sequential program that, initially, selects an initiator randomly out of the m that were chosen for the run, lets it fire, and queues up the messages it sends for reception by the destination nodes. Subsequently, after all initiators have had a chance to proceed in this way, a list is maintained containing all nodes with nonempty input-message queues. One of them is selected at random and the processing of its head-of-queue message is carried out (again with the possible queuing of messages for consumption by other nodes). This is repeated until all queues are empty. The directed graph S can be of one of types (i) -- (iii), as explained next. In order to keep the com- putational effort reasonably bounded with current technology, the directed graph D of which S is the GSCC has n = 100 nodes. (i) In this case D is sampled from the cortical model described above. As explained, the corresponding S is expected to have N ≈ 90 nodes. The expected in- or out-degree in D is 3.7. (ii) In this case D is sampled from the generalized Erdos-R´enyi model of directed graphs [61]. Given the expected in- or out-degree z, an edge is placed from each node i to each node j 6= i with probability z/(n − 1). For consistency with type-(i) graphs we use z = 3.7, in which case S is such that N ≈ 100 with high probability. (iii) In this case D has a deterministic structure and is by construction strongly connected. So S = D and N = 100. The structure of D is that of the directed circulant graph [62] with in- or out-degree equal to ⌈3.7⌉ = 4. If we number the n nodes 0, 1, . . . , n − 1, then in D every node i has nodes i + 1, i + 2, i + 3, and i + 4 (modulo n) as out-neighbors. 8 For each fixed S, 0.2N nodes are chosen uniformly at random to be inhibitory, so long as no two of these nodes are connected by an edge [note that, if S is of type (iii), then all 20 inhibitory nodes are in fact placed deterministically, since necessarily they are found at equal intervals as we traverse the nodes in the order 0, 1, . . . , n − 1]. Moreover, for each S node potentials and synaptic weights are chosen uniformly at random from the intervals [v0, vt] and [0, 1], respectively. For type-(iii) graphs this is the only source of nondeterminism. We use 50 S instances of each type. For each fixed S we use 50 000 run sequences of algorithm A, each sequence comprising 10 000 runs. The first run in a sequence starts from the node potentials and synaptic weights chosen for the graph. Each subsequent run in the sequence starts from the node potentials and synaptic weights at which the previous run ended. Along each sequence we observe the behavior of ρ− ij and ρ+ ij for each pair i, j of distinct nodes at six checkpoints. The first of these occurs before any run actually takes place. Each of the remaining five occurs after 2 000 additional runs have elapsed. The observation that takes place at a checkpoint is based on 100 additional runs, called side runs, each starting at its own set of m randomly chosen initiators and from the node potentials and synaptic weights that are current at the checkpoint. At the end of the side runs, the main sequence of runs is resumed from these same node potentials and synaptic weights. It is important to note that this computational setup would entail a considerable amount of processing even if the side runs were excluded, since for fixed S algorithm A would be run to completion 5×108 times. These runs are grouped into sequences so that the cumulative action of the model's weight-update rule can be effected, but we also need many independent sequences to account for the inherent nondeterminism of our model's asynchronous setting. As defined, however, our synchronization measures are properties of a single run, so they too need to be averaged out over many runs. We might have chosen to do so directly over the runs that correspond to checkpoints, that is, over 50 000 runs per checkpoint (one for each sequence). Each of these runs, however, starts at the set of node potentials and synaptic weights that are current in its sequence, so for each such set only one run would take place. Introducing side runs has been a means to increase this number to 100, with the consequence of elevating the overall number of runs for fixed S to 5.3 × 108. The contribution of each side run concerning a fixed pair i, j of distinct nodes is to record the values ij and ρ+ of ρ− ij at the end of the run, as well as tag them with the labels δmin = min{δij, δji} and δmax = max{δij, δji}, where δij and δji are the directed distances between the two nodes in S (from i to j and from j to i, respectively). After all 2.5 × 108 side runs for a graph type have been completed at a checkpoint, we calculate the average values of ρ− ij over all node pairs having the same tags. These averages are henceforth denoted by ρ− and ρ+, respectively. ij and ρ+ Our results are presented in Figures 2 and 3 for type-(i) graphs, Figures 4 and 5 for type-(ii) graphs, and Figures 6 and 7 for type-(iii) graphs. The former figure in each pair refers to ρ−, the latter figure to ρ+. Each figure comprises six panels, each panel for each of the six observational checkpoints. The A panels refer to the first checkpoints, the B panels to the second checkpoints, and so on. Each panel is organized as a two-dimensional array and gives the ρ− or ρ+ averages as a function of the node pairs' δmin values (as abscissas) and δmax values (as ordinates). We display these averages by means of a color code that assigns different colors to different intervals inside [0, 1] suitably. The hues we use vary from a dark shade of red to a dark shade of purple, indicating the lowest interval and the highest one, respectively. We note that this choice of colors is the same through all the figures and that the colors always correspond to the same intervals. Each average displayed in these figures refers to directed cycles in the S graphs whose length is δmin + δmax for the particular δmin and δmax values in question. Because of the strongly connected nature of S, every two nodes belong to at least one common directed cycle. By the definitions of δmin and δmax, each average plotted in the figures is therefore relative to all node pairs for which the shortest of these cycles has the same length. In particular, traversing any of the array diagonals for which δmin + δmax is a constant merely shifts the relative positions of the two nodes involved in each pair on the shortest 9 directed cycle that they share. We henceforth refer to the value of δmin + δmax for a certain node pair as the pairs' girth.3 Moreover, we refer to the pair as being more or less balanced on the directed cycle of length δmin + δmax, depending respectively on whether δmax − δmin is close to 0 or not (i.e., close to the δmin = δmax diagonal of the array). All panels in Figures 2 through 7 display their data inside the upper triangular region relative to the δmin = δmax diagonal of the array. All blank spots inside this region refer to δmin, δmax pairs that never occurred in any of the S instances we used. This can be verified by resorting to Figure 8, where the probability distributions for the occurrence of these pairs are shown for type-(i) and type-(ii) graphs (in panels A and B respectively, through the use of color codes similar to those of the previous figures). As for type-(iii) graphs, it follows easily from their definition that either δmin + δmax = N/4 or δmin + δmax = N/4 + 1, respectively 25 or 26 for N = 100, so in this case these fixed girth values are the constraints determining the appearance of blank spots (that is, they appear when the constraints are violated). Figure 8 is also useful in helping elucidate which of the various possible girth values in type-(i) and type-(ii) graphs are the most common. Readily, girths of about 18 or less are by far the most common in type-(i) graphs. This value becomes about 12 for type-(ii) graphs. In the forthcoming discussion, we use these rough delimiters to characterize what happens to most node pairs (i.e., those whose girth values are overwhelmingly the most common). Discussion The results for ρ−, given in Figures 2, 4, and 6, provide a clear picture of what is to be expected regarding the overall synchronization that may be present in the flow of messages as they get received at the neurons. In the case of type-(i) graphs (Figure 2), already at the first checkpoint most node pairs i, j have ρ− ij values in the interval (0.7, 0.85]. Four thousand runs later (that is, at the third checkpoint) this holds for the interval (0.9, 0.95]. At the sixth and last checkpoint, the interval is (0.9, 1]. A closely analogous conclusion holds in the case of type-(ii) graphs (Figure 4), now with the intervals (0.6, 0.85] for the first checkpoint, (0.9, 0.95] for the third, and (0.9, 1] for the last one. As for type-(iii) graphs (Figure 6), their rigidly constrained girth values lead ρ− ij to be concentrated inside the interval (0.7, 0.75] at the first checkpoint for nearly all node pairs. Similarly, already at the third checkpoint the situation that we observe in the last checkpoint has been established and ρ− ij is concentrated in the interval (0.9, 0.95] for nearly all node pairs. It is curious to observe for type-(iii) graphs that, at all checkpoints, all node pairs of girth 26 for which δmin = 1 have ρ− ij values occupying intervals one or two notches above the intervals we gave for nearly all pairs, that is, (0.8, 0.85] for the first checkpoint and (0.95, 1] for the third checkpoint and onwards. Revisiting Figures 2 and 4, we see that a similar conclusion holds also for type-(i) and type-(ii) graphs: Node pairs for which δmin is near 1 tend to have a slightly higher ρ− ij value if their girth is sufficiently large. The results for ρ+, which relates to how much synchronization may be present as neurons fire, tell a story that differs from that of ρ− in important ways. The first of these differences is clear from Figures 3, 5, and 7: For all graph types the range of occurring ρ+ values is wider by roughly 50 -- 100% than that of the ρ− values. In fact, the ρ+ values are now scattered inside the (0.4, 1] interval for all graph types at the earliest checkpoints and inside (0.5, 1] at the latest ones. Readily, then, according to our two measures of synchronization there appear to be substantially fewer synchronization possibilities in the firing of neurons than in the accumulation of potential as reflected by the reception of messages. This can be quantified by examining the data more closely, as follows. For type-(i) graphs (Figure 3), at the first checkpoint most node pairs have ρ+ ij values in the interval (0.4, 0.8]. At the last checkpoint, if we ignore all node pairs for which δmin = 1 for the time being then 3In a free extension, to a node pair, of the homonymous notion in graph theory that concerns the entire graph in the undirected case [63]. 10 nearly all node pairs have ρ+ ij values in the interval (0.9, 1] [the single exception is that of δmin = δmax = 2, for which the interval is (0.85, 0.9]]. The case of type-(ii) graphs (Figure 5), still disregarding all δmin = 1 entries, is closely analogous to that of type-(i) graphs, the differences being that now most node pairs span the larger interval (0.4, 1] already at the first checkpoint and that, without exception, at the last checkpoint all node pairs hit the interval (0.9, 1]. Finally, if we go on focusing on node pairs for which δmin > 1 exclusively, then for type-(iii) graphs (Figure 7) all node pairs have ρ+ ij values in the interval (0.45, 0.5] at the first checkpoint. At the last checkpoint, on the other hand, this interval becomes (0.85, 0.9] if δmin = 2, (0.9, 0.95] if δmin > 2. As Figures 3, 5, and 7 demonstrate, the δmin = 1 case sets itself apart from the others for all graph types at nearly all checkpoints. For example, if we concentrate on the last checkpoint, at which we believe the model's dynamics to have already settled into some sort of stationary regime [12], then for both type- (i) and type-(ii) graphs it holds that ρ+ ij tends to increase from some value in the interval (0.55, 0.6] when the girth is 2 to some value in the interval (0.9, 1] when the girth is close to the rough upper bound we set earlier for declaring most node pairs to have been counted [i.e., roughly 18 for type (i), roughly 12 for type (ii)]. Naturally, increasing the girth while δmin is held fixed at 1 implies considering node pairs that are progressively more imbalanced, since δmax grows with the girth. The case of type-(iii) graphs is different in this respect, as for δmin = 1 all node pairs have ρ+ ij values in the interval (0.6, 0.65], regardless of which value of δmax is in question (either 24 or 25). These pairs, however, are of course highly imbalanced as well. As we mentioned above, a notion that is becoming increasingly central to the study of synchronization in the brain is that of feedback loops, both in a temporal sense (as LFPs and neuronal firing patterns exert influence on one another) and in a spatial sense (as the LFPs of spatially separated areas affect one another). The results we have obtained with our algorithmic model of a cortex lend support to this notion both in the temporal and in the spatial sense. In the temporal sense we have found ample evidence that our distributed algorithm A is capable of promoting abundant opportunities both for potential to be accumulated in a synchronized way as messages are received at the neurons and for neurons to fire in a synchronized manner. We have found that this holds across all three graph types, from the cortical model first introduced in [12], to an Erdos- R´enyi directed graph, to the completely deterministic and tautly shaped structure of a directed circulant graph. In our view, this independence from the graph's structural characteristics points at an inherent ability of algorithm A at providing some of the elements that help give rise to brain synchronization. Notwithstanding this, our results also do shed some light on the role played by graph structure. As it happens, of the nondeterministic graph types only type-(i) graphs provide the opportunity of long-distance (in the sense of graph distances) synchronization in the two senses we have studied, since distances in type-(ii) graphs are significantly shorter. In the spatial sense there are two important issues to be highlighted. The first one is that, although by Figure 8 node pairs having higher-than-18 girth are rare, they do occur and have yielded high ρ− and ρ+ values at all the observational checkpoints. The exact significance this may have in the case of real cortices is unknown, to the best of our knowledge, since spatial feedback loops are known only for very small graph distances [2, 3]. So the fact that they may also occur at significantly larger graph distances remains a tantalizing possibility. The second important issue is the presence of such strong dependence of ρ+ on a node pair's girth when δmin = 1 as we observed. Our results indicate that in this case the synchronization of neuronal spikes is favored on feedback loops involving highly imbalanced pairs of neurons (i.e., node pairs for which δmax ≫ 1). As with the first issue, the significance this may have for real cortices is unknown and merits special attention as further data are obtained. All our results depend strongly on the measures of synchronization we gave in Equations (3) and (7). They also depend on the model summarized above, but that is now backed up by interesting validating finds [12, 13] and has therefore proven its usefulness as an artificial-life abstraction. It then seems that furthering our study of emerging synchronization properties depends on validating our two measures in 11 a way that ties our causality-based definitions to real data as tightly as possible. We expect to be able to do this as further insight into real cortices becomes available. Acknowledgments We acknowledge partial support from CNPq, CAPES, and a FAPERJ BBP grant. References 1. Nunez PL, Srinivasan R (2006) Electric Fields of the Brain: The Neurophysics of EEG. New York, NY: Oxford University Press. 2. Berens P, Logothetis NK, Tolias AS (2012) Local field potentials, BOLD and spiking activity: In: Kriegeskorte N, Kreiman G, editors, Visual Relationships and physiological mechanisms. Population Codes: Towards a Common Multivariate Framework for Cell Recording and Functional Imaging, Cambridge, MA: The MIT Press. pp. 599 -- 623. 3. Fries P (2005) A mechanism for cognitive dynamics: Neuronal communication through neuronal coherence. Trends Cogn Sci 9: 474 -- 480. 4. Canolty RT, Ganguly K, Kennerley SW, Cadieu CF, Koepsell K, et al. (2010) Oscillatory phase coupling coordinates anatomically dispersed functional cell assemblies. Proc Natl Acad Sci USA 107: 17356 -- 17361. 5. Canolty RT, Knight RT (2010) The functional role of cross-frequency coupling. Trends Cogn Sci 14: 506 -- 515. 6. Katzner S, Nauhaus I, Benucci A, Bonin V, Ringach DL, et al. (2009) Local origin of field potentials in visual cortex. Neuron 61: 35 -- 41. 7. Huerta R, Bazhenov M, Rabinovich MI (1998) Clusters of synchronization and bistability in lattices of chaotic neurons. Europhys Lett 43: 719 -- 724. 8. Lago-Fern´andez LF, Huerta R, Corbacho F, Siguenza JA (2000) Fast response and temporal co- herent oscillations in small-world networks. Phys Rev Lett 84: 2758 -- 2761. 9. Masuda N, Aihara K (2004) Global and local synchrony of coupled neurons in small-world networks. Biol Cybern 90: 302 -- 309. 10. Barbour B, Brunel N, Hakim V, Nadal JP (2007) What can we learn from synaptic weight distri- butions? Trends Neurosci 30: 622 -- 629. 11. Song S, Sjostrom PJ, Reigl M, Nelson S, Chklovskii DB (2005) Highly nonrandom features of synaptic connectivity in local cortical circuits. PLoS Biol 3: e68. 12. Nathan A, Barbosa VC (2010) Network algorithmics and the emergence of the cortical synaptic- weight distribution. Phys Rev E 81: 021916. 13. Nathan A, Barbosa VC (2011) Network algorithmics and the emergence of information integration in cortical models. Phys Rev E 84: 011904. 14. Balduzzi D, Tononi G (2008) Integrated information in discrete dynamical systems: Motivation and theoretical framework. PLoS Comput Biol 4: e1000091. 12 15. Barbosa VC (1996) An Introduction to Distributed Algorithms. Cambridge, MA: The MIT Press. 16. Fisher J, Henzinger TA (2007) Executable cell biology. Nat Biotechnol 25: 1239 -- 1249. 17. Fisher J, Harel D, Henzinger TA (2011) Biology as reactivity. Commun ACM 54(10): 72 -- 82. 18. Forbes N (2000) Life as it could be: Alife attempts to simulate evolution. IEEE Intell Syst 15(6): 2 -- 7. 19. Lindley D (2010) Brains and bytes. Commun ACM 53(9): 13 -- 15. 20. Sporns O, Chialvo DR, Kaiser M, Hilgetag CC (2004) Organization, development and function of complex brain networks. Trends Cogn Sci 8: 418 -- 425. 21. Sporns O, Tononi G, Kotter R (2005) The human connectome: A structural description of the human brain. PLoS Comput Biol 1: 245 -- 251. 22. Achard S, Salvador R, Whitcher B, Suckling J, Bullmore E (2006) A resilient, low-frequency, small- world human brain functional network with highly connected association cortical hubs. J Neurosci 26: 63 -- 72. 23. Bassett DS, Bullmore E (2006) Small-world brain networks. Neuroscientist 12: 512 -- 523. 24. He Y, Chen ZJ, Evans AC (2007) Small-world anatomical networks in the human brain revealed by cortical thickness from MRI. Cereb Cortex 17: 2407 -- 2419. 25. Honey CJ, Kotter R, Breakspear M, Sporns O (2007) Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proc Natl Acad Sci USA 104: 10240 -- 10245. 26. Reijneveld JC, Ponten SC, Berendse HW, Stam CJ (2007) The application of graph theoretical analysis to complex networks in the brain. Clin Neurophysiol 118: 2317 -- 2331. 27. Sporns O, Honey CJ, Kotter R (2007) Identification and classification of hubs in brain networks. PLoS ONE 2: e1049. 28. Stam CJ, Reijneveld JC (2007) Graph theoretical analysis of complex networks in the brain. Non- linear Biomed Phys 1: 3. 29. Yu S, Huang D, Singer W, Nikoli´c D (2008) A small world of neuronal synchrony. Cereb Cortex 18: 2891 -- 2901. 30. Chialvo DR (2010) Emergent complex neural dynamics. Nat Phys 6: 744 -- 750. 31. Modha DS, Ananthanarayanan R, Esser SK, Ndirango A, Sherbondy AJ, et al. (2011) Cognitive computing. Commun ACM 54(8): 62 -- 71. 32. Bornholdt S, Schuster HG, editors (2003) Handbook of Graphs and Networks. Weinheim, Germany: Wiley-VCH. 33. Newman M, Barab´asi AL, Watts DJ, editors (2006) The Structure and Dynamics of Networks. Princeton, NJ: Princeton University Press. 34. Bollob´as B, Kozma R, Mikl´os D, editors (2009) Handbook of Large-Scale Random Networks. Berlin, Germany: Springer. 35. Abeles M (1991) Corticonics: Neural Circuits of the Cerebral Cortex. Cambridge, UK: Cambridge University Press. 13 36. Ananthanarayanan R, Modha DS (2007) Anatomy of a cortical simulator. In: Proc. of the 2007 ACM/IEEE Conf. on Supercomputing. New York, NY: ACM, p. 3. 37. Ananthanarayanan R, Esser SK, Simon HD, Modha DS (2009) The cat is out of the bag: Cortical In: Proc. of the Conf. on High Performance simulations with 109 neurons and 1013 synapses. Computing Networking, Storage and Analysis. New York, NY: ACM, p. 63. 38. Dorogovtsev SN, Mendes JFF, Samukhin AN (2001) Giant strongly connected component of di- rected networks. Phys Rev E 64: 025101(R). 39. Newman MEJ (2005) Power laws, Pareto distributions and Zipf's law. Contemp Phys 46: 323 -- 351. 40. Egu´ıluz VM, Chialvo DR, Cecchi GA, Baliki M, Apkarian AV (2005) Scale-free brain functional networks. Phys Rev Lett 94: 018102. 41. van den Heuvel MP, Stam CJ, Boersma M, Hulshoff Pol HE (2008) Small-world and scale-free organization of voxel-based resting-state functional connectivity in the human brain. Neuroimage 43: 528 -- 539. 42. Modha DS, Singh R (2010) Network architecture of the long-distance pathways in the macaque brain. Proc Natl Acad Sci USA 107: 13485 -- 13490. 43. Bullmore E, Sporns O (2009) Complex brain networks: Graph theoretical analysis of structural and functional systems. Nat Rev Neurosci 10: 186 -- 198. 44. Erdos P, R´enyi A (1959) On random graphs. Publ Math (Debrecen) 6: 290 -- 297. 45. Freeman WJ, Kozma R, Bollob´as B, Riordan O (2009) Scale-free cortical planar networks. In: Bollob´as et al. [34], pp. 277 -- 324. 46. Sporns O (2011) Networks of the Brain. Cambridge, MA: The MIT Press. 47. Amaral LAN, Scala A, Barth´elemy M, Stanley HE (2000) Classes of small-world networks. Proc Natl Acad Sci USA 97: 11149 -- 11152. 48. Freeman WJ (2007) Scale-free neocortical dynamics. Scholarpedia 2(2): 1357. 49. Honey CJ, Sporns O, Cammoun L, Gigandet X, Thiran JP, et al. (2009) Predicting human resting- state functional connectivity from structural connectivity. Proc Natl Acad Sci USA 106: 2035 -- 2040. 50. Kaiser M, Hilgetag CC (2004) Modelling the development of cortical systems networks. Neuro- computing 58 -- 60: 297 -- 302. 51. Kaiser M, Hilgetag CC (2004) Spatial growth of real-world networks. Phys Rev E 69: 036103. 52. Abbott LF, Nelson SB (2000) Synaptic plasticity: Taming the beast. Nat Neurosci 3: 1178 -- 1183. 53. Song S, Miller KD, Abbott LF (2000) Competitive Hebbian learning through spike-timing- dependent synaptic plasticity. Nat Neurosci 3: 919 -- 926. 54. Bi GQ, Poo MM (1998) Synaptic modifications in cultured hippocampal neurons: Dependence on spike timing, synaptic strength, and postsynaptic cell type. J Neurosci 18: 10464 -- 10472. 55. Bi GQ, Poo MM (2001) Synaptic modification by correlated activity: Hebb's postulate revisited. Annu Rev Neurosci 24: 139 -- 166. 14 56. Kepecs A, van Rossum MCW (2002) Spike-timing-dependent plasticity: Common themes and divergent vistas. Biol Cybern 87: 446 -- 458. 57. Henriksen RN (2011) Practical Relativity: From First Principles to the Theory of Gravity. Chich- ester, UK: Wiley. 58. Lamport L (1978) Time, clocks, and the ordering of events in a distributed system. Commun ACM 21: 558 -- 565. 59. Innocenti GM, Lehmann P, Houzel JC (1994) Computational structure of visual callosal axons. Eur J Neurosci 6: 918 -- 935. 60. Salami M, Itami C, Tsumoto T, Kimura F (2003) Change of conduction velocity by regional myelination yields constant latency irrespective of distance between thalamus and cortex. Proc Natl Acad Sci USA 100: 6174 -- 6179. 61. Karp RM (1990) The transitive closure of a random digraph. Random Struct Algor 1: 73 -- 93. 62. Lonc Z, Parol K, Wojciechowski JM (2001) On the number of spanning trees in directed circulant graphs. Networks 37: 129 -- 133. 63. Bollob´as B (1998) Modern Graph Theory. New York, NY: Springer. 15 k1 k3 k2 k4 A B n o r u e N k1 k2 k3 k4 0 1 2 3 4 5 6 7 8 9 10 Depth Figure 1. Directed graphs S and P . A run of algorithm A on graph S (A) started by the single initiator k1 may result in the set of events shown in panel B as the node set of graph P . Events are shown in panel B either as filled circles, in the case of firing events, or as empty circles, otherwise. Each event is positioned against a background grid that highlights the node of S at which it occurred (this is given by the grid's rows) and its depth (the grid's columns). An edge of P either joins consecutive events occurring at the same node of S or corresponds to a message. We omit the former from panel B for clarity. An event's depth is the greatest number of message edges on a directed path arriving at it. Every event (except the first one at the initiator) has exactly one incoming message edge. If it is a firing event, then it also has as many outgoing message edges as the corresponding node in graph S has out-neighbors. Whenever two or more events occurring at the same node of S have the same depth, they are shown as close to the correct grid point as possible. 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 x a m A C E 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 d min 5 10 15 20 5 10 15 20 5 10 15 20 16 B D F 5 10 15 20 5 10 15 20 5 10 15 20 Figure 2. Average value of ρ− given for the first checkpoint (A) through the sixth (F). ij for type-(i) graphs as a function of δmin and δmax. Data are d 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 x a m A C E 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 d min 5 10 15 20 5 10 15 20 5 10 15 20 17 B D F 5 10 15 20 5 10 15 20 5 10 15 20 Figure 3. Average value of ρ+ given for the first checkpoint (A) through the sixth (F). ij for type-(i) graphs as a function of δmin and δmax. Data are d 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 x a m A C E 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 d min 5 10 15 20 5 10 15 20 5 10 15 20 18 B D F 5 10 15 20 5 10 15 20 5 10 15 20 Figure 4. Average value of ρ− given for the first checkpoint (A) through the sixth (F). ij for type-(ii) graphs as a function of δmin and δmax. Data are d 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 x a m A C E 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 d min 5 10 15 20 5 10 15 20 5 10 15 20 19 B D F 5 10 15 20 5 10 15 20 5 10 15 20 Figure 5. Average value of ρ+ given for the first checkpoint (A) through the sixth (F). ij for type-(ii) graphs as a function of δmin and δmax. Data are d 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 x a m A C E 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 0.4 < r − £ 0.45 0.45 < r − £ 0.5 0.5 < r − £ 0.55 0.55 < r − £ 0.6 0.6 < r − £ 0.65 0.65 < r − £ 0.7 0.7 < r − £ 0.75 0.75 < r − £ 0.8 0.8 < r − £ 0.85 0.85 < r − £ 0.9 0.9 < r − £ 0.95 0.95 < r − £ 1 d min 5 10 15 20 5 10 15 20 5 10 15 20 20 B D F 5 10 15 20 5 10 15 20 5 10 15 20 Figure 6. Average value of ρ− given for the first checkpoint (A) through the sixth (F). ij for type-(iii) graphs as a function of δmin and δmax. Data are d 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 x a m A C E 25 20 15 10 5 25 20 15 10 5 25 20 15 10 5 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 0.4 < r + £ 0.45 0.45 < r + £ 0.5 0.5 < r + £ 0.55 0.55 < r + £ 0.6 0.6 < r + £ 0.65 0.65 < r + £ 0.7 0.7 < r + £ 0.75 0.75 < r + £ 0.8 0.8 < r + £ 0.85 0.85 < r + £ 0.9 0.9 < r + £ 0.95 0.95 < r + £ 1 d min 5 10 15 20 5 10 15 20 5 10 15 20 21 B D F 5 10 15 20 5 10 15 20 5 10 15 20 Figure 7. Average value of ρ+ given for the first checkpoint (A) through the sixth (F). ij for type-(iii) graphs as a function of δmin and δmax. Data are d 22 B (0.0001, 0.0003] (0.0003, 0.001] (0.001, 0.003] (0.003, 0.01] (0.01, 0.03] (0.03, 0.1] 25 20 15 10 5 x a m A (0.0001, 0.0003] (0.0003, 0.001] (0.001, 0.003] (0.003, 0.01] (0.01, 0.03] (0.03, 0.1] 25 20 15 10 5 5 10 15 20 5 10 15 20 d min Figure 8. Probability distribution of δmin, δmax pairs for the type-(i) (A) and type-(ii) (B) graphs we used. d
1509.05556
1
1509
2015-09-18T09:16:12
Firing dynamics of an autaptic neuron
[ "q-bio.NC", "physics.bio-ph" ]
Autapses are synapses that connect a neuron to itself in the nervous system. Previously, both experimental and theoretical studies have demonstrated that autaptic connections in the nervous system have a significant physiological function. Autapses in nature provide self-delayed feedback, thus introducing an additional time scale to neuronal activities and causing many dynamic behaviors in neurons. Recently, theoretical studies have revealed that an autapse provides a control option for adjusting the response of a neuron: e.g., an autaptic connection can cause the electrical activities of the Hindmarsh-Rose neuron to switch between quiescent, periodic, and chaotic firing patterns; an autapse can enhance or suppress the mode-locking status of a neuron injected with sinusoidal current; and the firing frequency and interspike interval distributions of the response spike train can also be modified by the autapse. In this paper, we review recent studies that showed how an autapse affects the response of a single neuron.
q-bio.NC
q-bio
Firing dynamics of an autaptic neuron Hengtong Wang1 and Yong Chen2, ∗ 1School of Physics and Information Technology, Shaanxi Normal University, Xi'an 710119, China 2Center of soft matter physics and its application, Beihang University, Beijing 100191, China School of Physics and Nuclear Energy Engineering, Beihang University, Beijing 100191, China Autapses are synapses that connect a neuron to itself in the nervous system. Previously, both experimental and theoretical studies have demonstrated that autaptic connections in the nervous system have a significant physiological function. Autapses in nature provide self-delayed feedback, thus introducing an additional time scale to neuronal activities and causing many dynamic behaviors in neurons. Recently, theoretical studies have revealed that an autapse provides a control option for adjusting the response of a neuron: e.g., an autaptic connection can cause the electrical activities of the Hindmarsh-Rose neuron to switch between quiescent, periodic, and chaotic firing patterns; an autapse can enhance or suppress the mode-locking status of a neuron injected with sinusoidal current; and the firing frequency and interspike interval distributions of the response spike train can also be modified by the autapse. In this paper, we review recent studies that showed how an autapse affects the response of a single neuron. Keywords: Autapse, Self-delay feedback, single neuron, firing pattern PACS: 87.19.ll, 87.19.ls, 87.18.Sn, 87.19.lg I. INTRODUCTION The nervous system is a complex network of neurons, synapses, and other specialized cells, and through this system, neurons receive and transmit information be- tween different parts of the body [1]. The responses of neurons and the function of synapses have received con- siderable attention because of their status as building blocks of the nervous system [2 -- 4]. Nearly a century ago, neuroscientists found a special synaptic structure, the autapse, which is a synapse between different parts of the same neuron [5 -- 7]. Because of its odd structure, the self-synaptic connection has remained poorly under- stood. However, many recent studies found that autapses are much more widespread in the nervous system than previously thought. Autapses have been reported in var- ious brain areas, such as the neocortex, cerebellum, hip- pocampus, striatum, and substantia nigra [6, 8 -- 12]. In- terestingly, about 80% of cortical pyramidal neurons have autaptic connections [8]. Many studies have revealed that autapses are not merely curiosities, but play an authen- tic physiological role in the nervous system. Autapses can maintain persistent activity in the nervous system by mediating positive feedback [13]. Recent experiments also demonstrated that autapses are very important for the processing function of the brain [14]. Bacci et al. recorded the activity of fast spiking interneurons in acute brain slices of juvenile rats and found that autaptic transmission increased the spike- timing precision [15]. A somatic spike can evoke ex- citatory postsynaptic currents with various amplitudes through an autapse [11]. However, the function of au- tapses and their contribution to information processing ∗ [email protected] are still unclear [14, 16]. Thus, understanding the effect of autaptic activities on the responses of a neuron is a fundamental step to elucidating the process of informa- tion transfer in the nervous system [17 -- 20]. In fact, natural autapses are self-feedback connections in the nervous system and may allow a unique type of self-control, similar to the feedback in other systems [14]. Actually, feedback creates a circuit, or loop, that con- nects a system to itself and commonly occurs in many systems, e.g., gene regulatory networks [21, 22], popula- tion dynamics [23], and climate systems [24]. Feedback is also used extensively to control the state of a non- linear system, such as to stabilize periodic orbits or to control coherence resonance [25, 26]. Feedback is also very commonly used in the design of electric circuit el- ements [27]. In these systems, information about the past or the present influences the same phenomena in the present or future, respectively. Moreover, feedback has a marked effect on the dynamics of nonlinear systems. The frequency at the onset of the oscillation of a nonlinear system can be modified by the feedback loop [28]. Fur- thermore, self-delayed feedback with a fixed delay time can suppress chaos and also control steady states [29, 30]. In a simulation study, Popovych et al. showed that time- delayed feedback has the ability to desynchronize groups of model neurons [31]. A neuron with an autaptic connection can provide a distinctive physiological self-feedback model with many dynamic properties. Rusin et al. performed a time- delayed feedback stimulation of a group of cultured neu- rons and experimentally demonstrated that time-delayed feedback could cause synchronization of the action poten- tials of a group of neurons [32]. Prager et al. found that time-delayed feedback can facilitate the noise-induced os- cillation of a neuronal system [33]. Autaptic delayed feedback was found to modify the bursting in the dis- tribution of interspike intervals of a stochastic Hodgkin- Huxley (HH) neuron [34]. Such autaptic delayed feed- back can also reduce the spontaneous spiking activity of a stochastic neuron at the characteristic frequencies. The effect of an autapse on the spike rate of a single neuronal system shows dependence on the duration of the autapse activity [35]. Modulation of autaptic delay feedback can cause the dynamic behaviors of a Hindmarsh-Rose (HR) neuron to switch between quiescent, periodic, and chaotic firing patterns [36]. In this article, we review the current stat of knowledge of the effect of an autapse on single neurons. II. AUTAPTIC CONNECTION Scientists have long investigated the functions and properties of neurons and synapses. Neurons vary widely with different sizes and forms, and neuron responses are also involved in many dynamic behaviors. A synapse is a structure in the nervous system that permits a neuron (or nerve cell) to transmit an electrical or chemical signal to another cell (neural or otherwise) [37]. The self-synapse, or the autapse, was first described as a synapse between the axon of a pyramidal cell and its own dendrites by Van der Loos and Glaser in 1972 [5]. Before the report by Van der loos et al., other terms, such as 'self-excitation' and 'self-sensing', were used to describe the self-synaptic structure. For a long time, autapses in the nervous sys- tem seemed like anatomical curiosities with questionable functional significance. However, recent experiments be- gan to reveal how autapses might play an important role in brain function. Some reports also suggested that au- tapses are connected to some neural diseases [38]. Neu- robiology experiments have always focused on excitatory (glutamate-releasing) and inhibitory (GABA-releasing) autapses. To date, autaptic structures have been found in almost all parts of the nervous system [6]. Fig. 1 shows photos of the many autapses of a single rat hippocampal neuron. In nature, autaptic connections enable self-feedback of a neuron. Thus, in almost all current autapse studies, the model of a neuron with an autaptic connection con- tains a single compartment that exhibits a self-delayed feedback mechanism, as shown in Fig. 2. The delay time represents the elapsed time associated with the axonal propagation prior to the reintroduction of the signal into the neuron. The delay time is also believed to be one of the important properties of autaptic connections. Math- ematical models of an autapse in previous studies of the effect of self-synapses on neurons fall into two main cat- egories: one has linear self-coupling and describes a gap junction, and the other has nonlinear self-coupling and describes a chemical synapse, which may be an excitatory or inhibitory synapse. The specific mathematical form of the autapse depends on the chosen neuronal model. The autaptic connection of a simplified neuron model (such as the HR neuron, FitzHugh-Nagumo neuron, 2 FIG. 1. A single rat hippocampal neuron (arrowed, left) grown in culture on a 'microdot' of glia (flat gray cells), showing abundant autapses labeled with an antibody (small dark spots, right) [14]. Dendritic tree Axon Soma Autapse axonal tree FIG. 2. Schematic drawing of a neuron with autaptic feed- back. The structure in the circle represents an autaptic con- nection between dendritic and axonal terminals. Izhikevich neuron and others) can often be described by one of the following formulas [35, 36, 39]: • Linear coupling gives as electrical diffusive-type: Iaut = gaut[u(t − τ ) − u(t)]. (1) gaut is the autaptic conductivity. τ is the delay time. This form of autaptic current is proportional to the difference between the membrane potential at t and that at an earlier time point t − τ . • The chemical synapse function is modeled using the so-called fast threshold modulation (FTM) scheme [40, 41]: Iaut(t) = −gaut(u(t) − Vsyn)S(t − τ ), S(t − τ ) = 1/{1 + exp[−k(u(t − τ ) − θ)]}, (2) (3) where gaut is the autaptic intensity, and Vsyn is the synaptic reversal potential. Vsyn=2 and Vsyn=-2 correspond to excitatory and inhibitory autapses, respectively. For a conductance-based neuron model (such as the HH neuron, Morris-Lecar neuron, Connor-Stevens neu- ron and others), the mathematic model of the autapse is often given as follows [34, 35, 42, 43] • Linear coupling is also always set as electrical diffusive-type and has the same form as Eq. 1. • Conductance-based chemical autapses can be de- scribed using a bi-exponential function, which has been reported to fit well with experiment re- sults [44], Iaut(t) = G(t − tdelay)(V − Esyn), (4) where G(t − tdelay) is the autaptic conductance, and Esyn is the autaptic reversal potential. In this equation, we used Esyn = 0 mV for excitatory neu- rons and -80 mV for inhibitory neurons. The au- taptic conductance is defined as G(t) = gautσ(t − tf ire) σ(t) = exp(−t/td) − exp(−t/tr) td − tr (5) (6) with presynaptic spikes occurring at tf ire. The pa- rameters td and tr represent the decay and rise times of the function, respectively, and these pa- rameters determine the duration of the response. The autaptic rise and decay times were set to tr = 0.1 ms and td = 3 ms, respectively. In the following sections, we will review the two main forms of an autapse. Studies have shown that autapses that exhibit delayed feedback provide a control option for adjusting neuron responses. III. FIRING PATTERN TRANSITION IN A BURSTING NEURON Experimental observations have indicated that action potentials can occur with different firing patterns, and the primary firing patterns are spiking and bursting [45 -- 50]. Bursting is an extremely diverse general phe- nomenon in firing patterns exhibited by neurons in the central nervous system and spinal cord [51, 52]. Previ- ous studies on a bursting neuron with an autapse showed the novel dynamics and the transition of firing patterns induced by an autapse [36]. The simplified model of Hindmarsh and Rose has turned out to capture the features of experimentally mea- sured electrical data quite accurately, particularly for studies of the spiking-bursting behavior of neuron mem- brane potentials. Without an autapse, HR neurons exhibit many dy- namic behaviors, including quiescent, regular spiking, pe- riodic, and chaotic bursting firing patterns (see Fig. 1 in Ref. [36] and Ref. [53]). 3 (a) 2 0 u −2 5000 (d) 2 0 u 5200 t 5400 (b) 2 0 u −2 5000 (e) 2 0 u 5200 t 5400 (c) 2 0 u −2 5000 (f) 2 0 u 5200 t 5400 −2 5000 5200 t 5400 −2 5000 5200 t 5400 −2 5000 5200 t 5400 FIG. 3. Time courses of the membrane potentials in response to different Iext with gaut = 0.5 [36]. The blue curves repre- sent the time course of the HR neuron with an autapse, and the red curves indicate that without autapse. The presence of an autapse completely changes the fir- ing patterns of the original HR neuron. The firing pat- tern can be adjusted from a periodic or chaotic pattern to another periodic pattern or to a chaotic bursting pattern as the autaptic parameters change, independently of the original firing pattern. Fig. 3 shows the time courses of the membrane potentials of an HR neuron with an elec- trical autapse as an example. Importantly, the maximum action potential is increased by the electrical autapse. The ISI plot bifurcation diagrams of the HR neuron clearly show the transition between periodic and aperi- odic firing. For an electrical autapse, the periodic state transits to the chaotic state exhibiting an alternating be- havior as the time delay increased. With higher autap- tic intensity, this alternating behavior is more noticeable and occurs more frequently. For shorter delay times, the firing pattern of the HR neuron showed periodic spik- ing independent of the external DC input. The neuron with an excitatory chemical autapse exhibits chaotic fir- ing patterns in a larger area of gaut-τ space than that of the neuron with an electrical autapse. For strong exter- nal stimuli, the chaotic region of the combinational pa- rameters in the gaut-τ space is enlarged. The excitatory autapse plays a positive role in generating and enhancing chaos as a whole. As neurons with an inhibitory autaptic connection, the chaotic spiking of the HR neuron can be decreased and suppressed. With the proper inhibitory autaptic parameters, the HR neuron could be driven to a resting state. Fig. 4 shows the bifurcation of the ISIs of an HR neuron for the three types of autapses. Without an autapse, the firing pattern of an HR neu- ron switches from silent to periodic bursting and then to chaotic firing with period-doubling bifurcation as the DC current increased [54 -- 57]. For neurons with an autapse, the firing pattern can switch into any other firing pat- tern, regardless of the original pattern. As a whole, there are two main approaches to transition from a periodic to a chaotic firing pattern: discontinuous and continuous. The interspike interval return map is a useful approach to characterizing the transition to chaos. Fig. 5 shows an 4 1 + n I S I 80 60 40 20 0 0 80 60 40 20 0 0 (a) τ=2.182 20 40 60 80 (e) τ=41.97 20 40 60 80 80 60 40 20 0 0 80 60 40 20 0 0 (b) τ=2.183 20 40 60 80 (f) τ=41.99 (c) τ=2.184 20 40 60 80 (g) τ=42.00 80 60 40 20 0 0 80 60 40 20 20 40 60 80 0 0 ISI n 20 40 60 80 80 60 40 20 0 0 80 60 40 20 0 0 (d) τ=2.185 20 40 60 80 (h) τ=42.22 20 40 60 80 FIG. 5. An example of the two ways of transition (the dis- continuous transition [upper] and the continuous transition [bottom]) into chaos from periodic firing [36]. External stim- uli is set as Iext=2.67. 4 3 2 (a) t e a r g n i r i f l a n o (b) 15 t e a r g n i r i f l a n o 10 5 0 0 i t a r 1 i t a r 0 0 10 20 t delay 30 40 (c) 15 e t a r g n i r i f l a n o i t a r 10 5 0 0 (d) 15 e t a r g n i r i f l a n o i t a r 10 5 0 0 10 20 t delay 30 40 10 20 t delay 30 40 f=9 f=15 f=37 f=87 f=105 f=167 10 20 t delay 30 40 FIG. 6. Mode-locking of responses of an HH autaptic neuron with different delay times [42]. The lines with different mark- ers denote the firing rate of the neuron excited by different input frequencies. tem. Thus, the neuron-autapse system may exhibit very complex dynamics, due to the interplay between autaptic delayed feedback, an external periodic stimulus, and the intrinsic activity of the neuron. In previous studies, the model neuron-autapse system often contained an HH neuron and an autapse. The HH model is a conductance-based model that describes how action potentials in neurons are initiated and prop- agated [62]. In this section, we review the mode-locking firing pattern of an HH neuron with an electrical autapse that has the same mathematical form as Eq. (1). Without an autapse, the mode-locking behaviors de- pend on the values of the stimulus frequencies and am- plitudes. As shown in Fig. 1 in Ref [42], the Arnold tongues in the frequency and amplitude space show the overall features of various phase-locked states and provide the different p : q (denoting output action potentials per input spikes) mode-locking regions. Ref [58] also shows FIG. 4. Bifurcation diagram of the interspike interval of an HR neuron with an autapse versus the autaptic delay time [36]. From left to right, the panels show the results of the electrical autapse, excitatory chemical autapse, and in- hibitory chemical autapse. example of the two approaches for the transition of a neu- ron to chaos. In the discontinuous transition, the neuron has a periodic bursting firing pattern because the delay time is short. When the delay time increases, the system suddenly enters a chaotic firing state. In the continuous transition, the neuron first displays a periodic firing pat- tern. As the autaptic parameters change, the number of spikes in the single burst as well as the period of the periodic bursting increase almost continuously, and the neuron transitions into the chaotic state. IV. MODE-LOCKING BEHAVIOR OF A REGULAR SPIKING NEURON When excited by a periodic stimulus, neurons respond with various mode-locking firing patterns and quasi- periodic states [42, 58 -- 61]. When an autapse is present, the mode-locking firing of a neuron can be switched to another state, depending on the chosen autapse param- eters. An autapse provides self-feedback and contributes an additional time scale to the dynamic neuronal sys- the bifurcation mechanisms that create the boundaries of the complex mode-locking structure. The presence of an autapse substantially modifies the mode-locking patterns of the neuron. For the same si- nusoidal stimulus, the neuron with an autapse can fire more or fewer action potentials than that of a neuron without an autapse. The activities of the neuron can also be driven into sub-threshold oscillation. The electri- cal autapse displays a regulating function that controls the mode-locking firing of the neuron. The time courses of the membrane potentials of the neuron with and with- out an autapse are shown in Fig. 2 of Ref. [42]. With an autapse, the modification of the mode-locking firing pattern depends on the autaptic conductivity and the de- lay time. When the synaptic conductivity is small, the mode-locking p : q is similar to that without an autaptic connection. When the synaptic conductivity is large, the neuron displays very complex mode-locking firing. The p : q value of mode-locked firing increases from zero to a very large value when the frequency of the sinusoidal cur- rent increases from zero to the "threshold" frequency (the minimal input frequency for which the neuron fires an ac- tion potential given a fixed input amplitude). For further increases in the frequency of the sinusoidal current, the p : q value of mode-locking firing decreases. In the case of extremely strong autaptic conductivity [see Fig. 3 (d) in Ref. [42]], the autaptic activities completely disrupt the original mode-locking firing of the neuron without an autapse. For continuous changes in the delay time, the model- locked firing of the neuron is also quite interesting. Fig. 6 shows the dependence of the mode-locking state on the delay time for different input frequencies and au- taptic conductivities. For weak autaptic conductivities [Fig. 6(a)], the p : q values of mode-locking are similar to those without self-feedback for the short delay time. As the delay time increases, the values of p : q mode-locking begin to fluctuate. The smaller the input frequency is, the larger the resulting fluctuation. For strong autap- tic intensities [shown in Fig. 6(b-d)], the p : q values of mode-locking decrease with increasing delay time in a stepwise manner. With further increases in the delay time, the p : q values of mode-locking suddenly jump to a very large value (larger than that without autaptic self- feedback) and then decrease smoothly. As the delay time increases, the p : q value of mode-locking of a neuron with a high autaptic intensity exhibits periodic behavior. With autaptic self-feedback, the responses of an HH neuron to a sinusoidal stimulus can have higher or lower p : q mode-locking responses than that of a neuron with- out an autapse. These dynamical behaviors depend on the autaptic intensity and the delay time. Moreover, the presence of an autapse increases the range of values for which the HH neuron spiking is locked with the sinusoidal current. When the autaptic intensity is weak, the mode- locking behaviors show no marked changes. That is, an autapse with a weak intensity will have a negligible ef- fect on the response of the neuron. For stronger autaptic 5 120 110 100 y c n e u q e r f g n i r i f 90 80 70 60 10 20 t delay 30 (ms) 40 50 120 (a) 1 ) 2 m c / S m 0.8 0.6 ( g t u a 0.4 0.2 0 (c) 1 0.8 0.6 ) 2 m c / S m ( g t u a 0.4 y c n e u q e r f g n i r i f 110 100 90 80 70 60 0.2 0 10 20 t delay 30 (ms) 40 50 FIG. 7. (Color online) Firing frequencies ( (a),(c)) and ISI distribution ((b), (d)) of a neuron with a chemical autapse (upper panels: excitatory autapse; lower panels: inhibitory autapse) excited by a DC current [43]. Dotted lines gives the odd multiples of half the intrinsic period of a single HH neuron excited by the corresponding DC current. intensities, the p : q value of the mode-locking increases and then decreases as the delay time increases, exhibiting nearly periodic behaviors. Thus, for sufficiently strong autaptic intensity, changing the delay time provides bet- ter regulation of mode-locking than does changing the autaptic intensity. The autaptic connection may also pro- vide a control option for adjusting mode-locked firing in a neural information process. V. RESPONSES OF A REGULAR SPIKING NEURON A. Response to DC currents When injected with a DC current, an HH neuron with- out an autapse displays regular spiking when the current intensity is larger than a critical current value. As for most neuronal models, increases in the current will in- crease the firing rate of the neuron, and the HH neu- ron undergoes a Hopf bifurcation. In the presence of an autapse, the output frequency can be higher or lower than the intrinsic frequency (the firing frequency of an HH neuron without an autapse) when the autaptic con- ductance and delay time are changed. Interestingly, the output frequency, as well as the output ISI distribution, shows periodic behaviors as the autaptic delay time in- creases. This periodicity of the changes in the output frequency (ISI distributions) in response to changes in the delay time is also similar to the intrinsic period of the neuron. When an HH neuron is connected to an elec- trical or excitatory chemical autapse, the firing rate and the ISI distribution of the response spike trains can both be either amplified or depressed. When the delay time approaches the odd multiples of the intrinsic half period of the firing of an HH neuron with an electrical or excita- tory chemical autapse, the ISIs of the spike train are very different from the intrinsic ISIs. When the delay time approaches a multiple of the intrinsic period, the ISIs of the spike train are similar to the intrinsic ISIs. However, the inhibitory chemical autapse can only suppress the firing responses of the neuron. The ISIs of the response spike train of a neuron with an inhibitory autapse are al- ways greater than the intrinsic period, whereas the corre- sponding response frequency is not less than the intrinsic response frequency. Current neurobiological experiments also reveal that autapses in the nervous system can be ei- ther self-inhibitory or self-excitatory, depending on their location on the neuron [6, 63 -- 65]. Because autaptic pulses perturb neural spiking through a process that is similar to that of an oscillating system with a fixed-delay feedback, the phase-response curve (PRC) theory can provide useful insight into the phenomena of the neuron-autapse system [66 -- 69]. For an oscillating system, the PRC describes the phase shift that occurs in response to a brief external stimulus [70, 71]. Fig. 8 gives the phase-response curves of the three types of autapses. The PRCs of a neuron with an electrical autapse or an excitatory chemical autapse can be nega- tive or positive, depending on the stimulus phase (delay time). Thus, an electrical autapse or excitatory chemical autapse could delay or advance the next spike. Therefore, the firing rate of a neuron with an electrical autapse or an excitatory chemical autapse can be higher or smaller than that of the neuron without an autapse. For the in- hibitory chemical autapse, however, the PRC is always positive, because the delay time increased. That is, an in- hibitory autapse always postpones the next spike. Thus, the firing rate of an HH neuron with an inhibitory au- tapse is not larger than that of the neuron without an autapse. When the DC current increases, the HH neuron with- out an autapse undergoes a Hopf bifurcation from a quiescent state to a periodic spiking state. Although the limit cycle and the oscillation period of the neuron have been disturbed by the autaptic current, the neuron- autapse system still undergoes the Hopf bifurcation, gen- erating a stable periodic orbit. Interestingly, an HH neu- ron with some special autaptic parameters does not fire regular action potentials and is attracted to the quies- cent state. This result reveals the phenomenon of spiking death that is induced by an autaptic connection, which is more clearly shown in the membrane potential and autap- tic current traces in Fig. 9. With an upper-threshold DC current, the neuron fires action potentials at first. After the delay time, the subsequent autaptic pulse reaches the neuron and drives the neuron to a fixed point. Moreover, the spiking death induced by the autapse is independent of the initial conditions of the neuron. As shown in the inset of Fig. 9 (a), the neuron-autapse system can return (a) e v r u c e s n o p s e r e s a h P 0.5 0 −0.5 0 0.5 phase (c) 0.05 0 e v r u c e s n o p s e r e s a h P −0.05 −0.1 −0.15 g aut g aut g aut g aut g aut =0.1 =0.3 =0.5 =0.7 =0.9 −0.2 0 0.5 phase (b) 0.2 0.1 0 e v r u c e s n o p s e r e s a h P 1 −0.1 0 0.5 phase 6 1 1 FIG. 8. (Color online) The phase-response curve of three types of autapses with different autaptic conductivity gaut, (a) electrical autapse, (b) excitatory chemical autapse, and (c) inhibitory chemical autapse. The external DC stimuli is set as 26 µA/cm2. (a) 40 0 −40 −80 ) v m ( V 490 320 150 −20 0 100 200 300 400 500 (b) 1 m 0.8 0.6 0.4 0.2 0 −100 time (ms) (c) 40 0 −40 −80 ) v m ( V (d) 0.18 490 320 0.16 m 150 −20 0.14 0.12 −50 0 V (mv) 50 0 10 20 30 time (ms) 40 50 −58 −56 V (mv) −54 FIG. 9. (Color online) (a) Traces of the membrane poten- tial (dashed lines) and autaptic current (solid lines) [43]. (b) The phase portrait corresponds to (a). (c) and (d) are the enlarged plots of (a) and (b), respectively. The different col- ors indicate different integral initial conditions (The specific set of parameters can be found in Ref. [43]). The inset in (a): The membrane potential of HH neuron-autapse system perturbed by a brief step pause (black solid line). to the fixed point even when the input is a brief exter- nal perturbation. The figure also shows that the spiking death induced by the autapse is stable. The Ref. [72] also shows a similar phenomenon: an HH neuron transitions from a limit cycle to a fixed point when the neuron is perturbed by an excitatory chemical synaptic pulse. B. Response to a random synaptic pulse-like input The assumed α form of the postsynaptic current model that is used in many works is perfect for generating pulse- like currents that are similar to the synaptic pulses ob- served experimentally and also enables easy modifica- tion of the ISI of the input current to investigate the input-output properties conveniently [60]. Injecting this type of synaptic pulse-like signal with different ISIs into the HH neuron, Hideo et al. have reported that the mean firing frequency decreases as the mean input ISI increases [73, 74]. When a neuron is injected with a synaptic-like pulsed current with random ISIs, the neu- ron displays interesting autapse-induced response behav- iors [43]. As the delay time increases, the frequency of the neuron with a sufficient electrical or excitatory chem- ical autapse shows nearly periodic behaviors. Such peri- odicity is not changed when the mean ISIs of the input synaptic current changes. When given the input synaptic pulses with a large mean value of ISIs, however, the re- sponse frequency of a neuron with an inhibitory autapse does not show periodic behavior when the autaptic delay time increases. Autaptic activities also influence the detailed response of the ISI distribution. For a short autaptic delay time, the output ISIs are distributed in an area [blue bars in Fig. 10] that is smaller than the input area (red lines in Fig. 10). When the delay time increases, the region in which the output ISIs are distributed does not change much compared with that for the neuron without an au- tapse (green lines in Fig. 10). For further increases in the delay time, the distribution of the output ISIs suddenly decreases almost to a single point (tdelay = 8.5 ms in Fig. 10). For these autaptic parameters, the responding spike train is strongly regulated. Then, the size of the distribution of ISIs increases slowly as the delay time is increased further. When the size of the distribution of the output ISIs in- creases to its maximum (this maximum distribution area is smaller than the area for the low delay time condition tdelay < 5.5ms), the distribution suddenly shrinks nearly to a point again (see tdelay = 20.5 ms in Fig. 10). For the entire range of delay times, the distribution of the output ISIs periodically exhibits the above behaviors when the delay time increases. Thus, the delayed feedback activi- ties of an autapse can act as a regulator that adjusts the ISIs of the output spike train. For some specific autaptic parameters, the resulting spike train of a neuron can be modified to obtain an almost regular spiking, even for a very random ISI input. Without an autapse, the neuron will filter the spikes with a short ISI and thus show low-pass filtering behav- ior when the neuron is injected with a random synaptic pulse-like current [73]. For a neuron with an autapse, long ISI pulses can be filtered in addition to the short ISI spikes. Thus, the neuron-autapse system displays com- plex filtering behaviors, including low-pass filtering and band-filtering behaviors. 7 t delay =1.5 1 0.5 t delay =3.5 1 0.5 t delay =5.5 0 0 1 20 40 t delay =8.5 20 40 t delay =13.5 0 0 1 20 40 t delay =17.5 0.5 0.5 20 40 t delay =20.5 0 0 1 0.5 20 0 0 40 20 ISI (ms) 0 0 1 20 40 t delay =24.5 20 40 t delay =28.5 0.5 0 0 40 20 40 ) t i n u . b r a ( s m a r g o t s H i 1 0.5 0 0 1 0.5 0 0 1 0.5 0 0 FIG. 10. (Color online) Histograms of output (blue bars) ISIs of a neuron with an electrical autapse [43]. The green lines gives the output ISI of a neuron without an autapse and the red lines give the ISI of the random input pulse train. When the delay time is short, an HH neuron with an electrical or excitatory chemical autapse acts as a low- pass filter similar to a neuron without an autapse and removes the spikes with short ISIs. When the autap- tic delay time is long enough, the neuron displays band- pass filtering behavior and removes both the short- and long-ISI spikes. The cut-off value for the ISIs (either the minimum or the maximum ISIs that will be retained in the output spike train) can be changed by changing the autaptic parameters. For some specific autaptic param- eters, the neuron can filter most of the input pulse, and the output spike train can be altered into a nearly regular spike train, even for an input with highly random ISIs. More interestingly, a neuron with an inhibitory chemi- cal autapse can only act as a low-pass filter. Thus, the inhibitory chemical autapse does not have a significant effect on the frequency of the output spike train. It can be conjectured that the filtering properties de- pend on the intrinsic properties of both the neuron and the autapse. For electrical or excitatory chemical au- tapses, the delay time will act as a border or as a cut-off value for the ISI filtering of synaptic pulses with a long ISI. The intrinsic filtering property of neurons removes the short input ISI pulses, and the time delay of the elec- trical or excitatory chemical autapse removes the long input ISI pulses. However, an inhibitory autapse delays the spikes of a neuron and thus filters the spikes that have short ISI pulses. Thus, the firing frequency of a neuron with an inhibitory autapse is not higher than that of a neuron without an autapse. VI. EFFECT OF NOISE ON THE FIRING DYNAMICS OF A NEURON capability of the HH neuron is strongly dependent on the cell size and autaptic strength. 8 Neuronal noise is random electrical fluctuations within neuronal networks and affects the patterns of neural ac- tivity in a determinant way [75 -- 78]. In the presence of noise, autapses can also substantially affect the response of a neuron. In this section, we review previous studies on the interplay of autapses and noise on the firing of an autaptic neuron. Considering the subthreshold dynamics of a neuron with interaction between autaptic-delayed feedback and noise, Masoller et al. investigated the firing patterns of an HH model of a thermoreceptor neuron with an elec- trical autaptic feedback in the presence of a Gaussian white noise [79]. In their studies, the neuron displays only subthreshold oscillations in the absence of feedback and noise. Their results show that the interaction among weak autaptic feedback, noise, and the subthreshold in- trinsic activity is nontrivial. The subthreshold oscillation amplitude can be enhanced by the autapse in the pres- ence of external noise, and this enhancement is more pro- nounced for certain delay values. For negative autaptic delay feedback, the firing rate can be lower than that of the noise-free situation, depending on the delay. This is because noise inhibits feedback-induced spikes by driving the neuronal oscillations away from the firing threshold. For positive autaptic delay feedback, there are regions of delay values where the noise-induced spikes are inhibited by the feedback; in this case, the autaptic feedback drives the neuronal oscillations away from the threshold. Li and his co-workers analyzed the effects of electri- cal autapses on the spiking dynamics of a stochastic HH neuron [34], considering the stochastic gating of ion chan- nels or the so-called intrinsic channel noise is considered. They found that the delayed feedback manifests itself in the occurrence of bursting and a rich multimodal in- terspike interval distribution, exhibiting a delay-induced reduction in the spontaneous spiking activity at charac- teristic frequencies. For small numbers of ion channels, the channel noise is sizable and the excitatory dynam- ics remain practically unaffected by the delay. However, smaller noise levels and stronger autaptic intensities in- duce different synchronization phenomena between the delay time and the intrinsic time scales. The delay time and the intrinsic time scales determine the number of spikes that will be induced and become subsequently locked during one delay epoch. Recently, Yilmaz and Ozer showed that the electrical autaptic delayed feedback either enhances or suppresses the weak signal detection, depending on the parameters of autapse and channel noise [80]. When the delay time is close to integer multiples of the period of the intrinsic oscillations, the autapse enhances the weak periodic sig- nal detection for the optimal values of the intrinsic noise and the autaptic intensity. The system response also ex- hibits stochastic resonance behavior, depending on the autaptic intensity. Moreover, the weak signal detection VII. CONCLUSION AND OUTLOOK Self-delayed feedback significantly affects nonlinear dy- namic systems because such self-feedback loops introduce a new time scale into the dynamics of the system [81, 82]. The system of a neuron with an autapse contains two time scales: the intrinsic time scale of the neuron and the time scale of the autapse. However, the situation is more complex in the nervous system, because there are many more time scales, including the time scales of the neurons, the synapse, the environment, and the autaptic connections. This complexity raises the question of how the activity time of an autapse influences the activities of coupled neurons in the nervous system. Autapses provide self-feedback circuits that are com- mon in the nervous system. Since the naming of au- tapses, many experimental studies have revealed that au- tapses play important roles in brain function [6, 11, 14, 15, 17, 18, 20]. The current theoretical studies on the topic of autaptic connections also reported the impor- tance of autapses and the many novel dynamic behaviors induced by the autapse [32, 34 -- 36, 42, 43]. Autapses provide a control option that can sufficiently adjust the firing behaviors of a neuron for any form of input stimu- lus, regardless of the neuron type. Autapses offer a new mechanism for switching between quiescent, periodic and chaotic firing patterns in bursting neurons [36]. Additionally, the nervous system responds rapidly to an external stimulus based on the transition between neuron firing patterns [40, 83]. Thus, the results of the firing pattern transition induced by an autapse also indicate that an autapse could act as an efficient tool for controlling the transition among different relevant neu- ronal activities in the nervous system. For an HH neuron with an autapse, the firing fre- quency and interspike interval distributions of the out- put spike train show periodic behavior when the delay time is increased [43]. When specific autaptic parameters are chosen, the response spike trains are nearly regular, and the ISI distribution covers a small area, even with a highly random input. This phenomenon emphassises the autapse-induced filtering behaviors of the neuron. These results about the autapse are useful for studying the con- trol of a nonlinear system. In neuronal systems, an autapse can take the form of recurrent excitation, i.e., the discharge of a neuron, pos- sibly after passing through the axon, can subsequently induce an excitatory response in the same neuron. This self-excitation mechanism is important for maintaining persistent activity, particularly the feeding behavior in sea slugs. Leonel et al. also reported that such processes play functional roles in amplifying activity in neuronal assemblies, causing reverberating activity, inducing some form of memory, or generating rhythmic patterns, such as those in central pattern generators [84]. The previous experimental data also indicate that the brain tissue ex- presses a novel form of self inhibition, namely autaptic inhibitory transmission in Fast-Spike (FS) cells slow self activities in interneurons [85]. Ma et al. also investi- gated the activities of a two-dimensional neural network containing autapses with different time delays and found that the autapses induced many novel collective behav- iors [86]. The effect of autapse on the synchronization of neural network have also been conducted in a group of HH neurons with small world network structure [87]. It was found that the neurons exhibit synchronization transitions as autaptic delay feedback is varied, and fine synchronized network activities occur when an optimal autaptic strength is chosen. Alberto Bacci and his col- leagues studied the autaptic self-inhibition of basket cells and the role of the autaptic connections of disinhibition within cortical circuits and argued that autaptic feed- back could have a dual function in temporally coordi- nating parvalbumin basket cells during cortical network 9 activity [88]. Another study also indicated that fast spik- ing neurons with autapses showed the strongest asyn- chronous release in brain slices obtained from patients with intractable epilepsy, and that such discharges may be involved in generating and regulating network activi- ties, including epileptic activity [38]. Although there are many studies on the topic of the autapse, the precise function of autapses and their con- tribution to information processing are remain unclear. There are also many open questions both in the exper- imental and theoretical areas: 1) What is the role of autapses in coordinating network activities? i.e., FS cell autapses in adjusting fast network synchrony. 2) What are the molecular mechanisms underlying autaptic asyn- chronous release and what is its relevance during physio- logical and pathological network activities? 3) What are the functions of autapses in the information propagation in the neural circuit and neural network? Addressing these questions will also facilitate our understanding of the fundamental mechanisms governing several core func- tions of cortical activities. [1] Kandel E R, Schwartz J H, Jessell T M, Siegelbaum S A and Hudspeth A J 2000 Principles of Neural Science (New York: McGraw-Hill Medical) [2] Bartos M, Vida I and Jonas P 2007 Nat. Rev. Neurosci. [23] Cabrera J L and Milton J G 2002 Phys. Rev. Lett. 89 158702 [24] Bonan G B 2008 Science 320 1444 [25] Postlethwaite C M and Silber M 2007 Phys. Rev. E 76 8 45 056214 [3] Bennett M V L and Zukin R S 2004 Neuron 41 495 [4] Connors B W and Long M A 2004 Annu. Rev. Neurosci. 27 393 [5] Van der Loos H and Glaser E M 1972 Brain Res. 48 355 [6] Bekkers J M 1998 Curr. Biol. 8 R52 [7] Yamaguchi K 2008 Autapse Encyclopedia of Neuroscience (M. D. Binder ed.), N Hirokawa and U Windhorst (Springer Berlin Heidelberg) pp 229-32 [26] Sethia G C, Kurths J, and Sen A 2007 Phys. Lett. A 364 227 [27] Chen W-K 2005 Circuit Analysis and Feedback Amplifier Theory (Bocan Raton, FL: CRC Press) [28] Gaudreault M, Drolet F and Vials J 2012 Phys. Rev. E 85 056214 [29] Ahlborn A and Parlitz U 2004 Phys. Rev. Lett. 93 264101 [30] Balanov A G, Janson N B and Scholl E 2005 Phys. Rev. [8] Lubke J, Markram H, Frotscher M and Sakmann B 1996 E 71 016222 J. Neurosci. 16 3209 [31] Popovych O V, Hauptmann C and Tass P A 2006 Biol. [9] Flight M H 2009 Nat. Rev. Neurosci. 10 316 Cybern. 95 69 [10] Branco T and Staras K 2009 Nat. Rev. Neurosci. 10 373 [11] Kimura F, Otsu Y and Tsumoto T 1997 J. Neurophysiol. [32] Rusin C G, Johnson S E, Kapur J and Hudson J L 2011 Phys. Rev. E 84 066202 77 2805 [33] Prager T, Lerch H P, Schimansky-Geier L and Scholl E [12] Tam´as G, Buhl E H and Somogyi P 1997 J. Neurosci. 17 2007 J. Phys. A 40 11045 6352 [34] Li Y, Schmid G, Hnggi P and Schimansky-Geier L 2010 [13] Saada R, Miller N, Hurwitz I and Susswein A J 2009 Phys. Rev. E 82 061907 Curr. Biol. 19 479 [35] Hashemi M, Valizadeh A and Azizi Y 2012 Phys. Rev. E [14] Ikeda K and Bekkers J M 2006 Curr. Biol. 16 R308 [15] Bacci A and Huguenard J R 2006 Neuron 49 119 [16] Bekkers J M 2009 Curr. Biol. 19, R296 [17] Salinas E and Sejnowski T J 2001 Nat. Rev. Neurosci. 2 539 85 021917 [36] Wang H, Ma J, Chen Y and Chen Y 2014 Commun. Nonlinear Sci. Numer. Simul. 19 3242 [37] Schacter D L, Gilbert D T, Wegner D M and Nock M K 2014 Psychology (New York, NY: Worth Publishers) [18] Chen Y, Yu L and Qin S M 2008 Phys. Rev. E 78 051909 [19] Chen Y, Zhang H, Wang H, Yu L and Chen Y 2013 PLoS [38] Jiang M, Zhu J, Liu Y, Yang M, Tian C, Jiang S, Wang Y, Guo H, Wang K and Shu Y 2012 PLoS Biol. 10 e1001324 ONE 8 e56822 [20] Quiroga R Q and Panzeri S 2009 Nat. Rev. Neurosci. 10 [39] Jun Ma H Q 2015 Int. J. of Mod. Phys. B 29 1450239 [40] Belykh I, de Lange E and Hasler M 2005 Phys. Rev. Lett. 173 94 188101 [21] Davidson E. and Levin M. 2005 Proc. Natl. Acad. Sci. [41] Buri´c N, Todorovi´c K and Vasovi´c N 2008 Phys. Rev. E U.S.A. 102 4935 78 036211 [22] Zhang H, Chen Y and Chen Y 2012 PLoS ONE 7 e51840 [42] Wang H, Sun Y, Li Y and Chen Y 2014 J. Theor. Biol. 10 358 25 [43] Wang H, Wang L, Chen Y and Chen Y 2014 Chaos 24 033122 [44] Connelly W M and Lees G 2010 J. Physiol. 588, 2047 [45] Cocatre-Zilgien JH, Delcomyn F 1992 J. Neurosci. Meth. Lett. A 125, 119 [67] Glass L and Zeng W-Z 1990 Ann.NY Acad.Sci. 591 316 [68] Lewis J E, Glass L, Bachoo M, and Polosa C 1992 J. Theor. Biol. 159 491 [69] Kunysz A M, Shrier A, and Glass L 1997 Am. J. Physiol. 41 19 273 C331 [46] Mainen Z F and Sejnowski T J 1996 Nature 382 363 [47] Izhikevich E M 2000 Int. J. Bifurcat. Chaos 10 1171 [48] Kepecs A, Wang X-J and Lisman J 2002 J. Neurosci. 22 9053 [49] Gu Hua-Guang,Zhu Zhou,Jia Bing. 2011 Acta Physica Sinica 60 100505 [70] Smeal R M, Ermentrout G B, and White J A 2010 Trans. R. Soc. B 365 2407 [71] Krogh-Madsen T, Butera R, Ermentrout G B, and Glass L, in Phase Response Curves in Neuroscience, edited by N.W. Schultheiss, Prinz A A., and Butera R J (Springer New York, 2012), pp. 33 [50] Yu H-T, Wang J, Deng B and Wei X-L 2013 Chin. Phys. [72] Wang S-J, Xu X-J, Wu Z-X, Huang Z-G, and Wang Y-H B 22 018701 [51] Wagenaar D A, Pine J and Potter S M 2006 BMC Neu- rosci. 7 11 [52] Yu H, Wang J, Deng B, Wei X, Wong Y K, Chan W L, 2008 Phys. Rev. E 78 061906 [73] Hasegawa H 2000 Phys. Rev. E 61 718 [74] Borkowski L S Phys. Rev. E 80 051914 [75] White J A, Rubinstein J T and Kay A R 2000 Trends in Tsang K M, Ziqi Y. et al. 2011 Chaos 21 013127 Neurosci. 23 131 [53] Tang G, Xu K and Jiang L 2011 Phys. Rev. E 84 046207 [54] Wang X-J 1993 Physica D 62 263 [55] Jm G-M 2003 Chaos 13 845 [56] Innocenti G, Morelli A, Genesio R and Torcini A 2007 [76] Jacobson G A, et al. 2005 J. Physiol. 564 145 [77] Wang H, Yu L and Chen Y 2009 Acta Phys. Sin. 58 5070 [78] Destexhe A and Rudolph-Lilith M 2012 Neuronal Noise (New York: Springer) Chaos 17 043128 [79] Masoller C, Torrent M C and Garcia-Ojalvo J 2008 Phys. [57] Innocenti G and Genesio R 2009 Chaos 19 023124 [58] Lee S-G and Kim S 2006 Phys. Rev. E 73 041924 [59] Che Y-Q, Wang J, Si W-J and Fei X-Y 2009 Chaos, Soli- tons & Fractals 39 454 Rev. E 78 041907 [80] Yilmaz E and Ozer M 2015 Physica A 421 455 [81] Soriano M C, Garcia-Ojalvo J, Mirasso C R and Fischer I 2013 Rev. Mod. Phys. 85 421 [60] Wang H, Wang L, Yu L and Chen Y 2011 Phys. Rev. E. [82] Park J-H, Huh S-H, Kim S-H, Seo S-J and Park G-T 83 021915 [61] Fellous J-M, Houweling, A R, Modi, R H, Rao, R P N, Tiesinga P H E and Sejnowski T J2001 J Neurophysiol. 85 1782 2005 IEEE Trans. Neural. Netw. 16 414 [83] Erichsen R J, Brunnet L G. 2008 Phys. Rev. E 78 061917 [84] G´omez L, Budelli R and Pakdaman K 2001 Phys. Rev. E 64 061910 [62] Hodgkin A L and Huxley A F 1952 The J. of Physiol. [85] Pawelzik H, Hughes D I and Thomson A M 2003 J Phys- 117 500 iol 546 701 [63] Bekkers J M 2003 Curr. Biol. 13, R433 [64] Gulledge A T and Stuart G J 2003 Neuron 37, 299 [65] Saada-Madar R, Miller N and Susswein A J 2012 J. Mol. Histol. 43 431 [86] Qin H, Wu Y, Wang C and Ma J 2015 Commun. Non- linear. Sci. Numer. Simulat. 23 164 [87] Wu Y, Gong Y and Wang Q 2015 Chaos 25 043113 [88] Deleuze C, Pazienti A and Bacci A 2014 Curr. Opin. [66] Lewis J, Bachoo M, Glass L, and Polosa C 1987 Phys. Neurobiol. 26 64
1602.08881
1
1602
2016-02-29T09:35:07
Brain State Control by Closed-Loop Environmental Feedback
[ "q-bio.NC" ]
Brain state regulates sensory processing and motor control for adaptive behavior. Internal mechanisms of brain state control are well studied, but the role of external modulation from the environment is not well understood. Here, we examined the role of closed-loop environmental (CLE) feedback, in comparison to open-loop sensory input, on brain state and behavior in diverse vertebrate systems. In fictively swimming zebrafish, CLE feedback for optomotor stability controlled brain state by reducing coherent neuronal activity. The role of CLE feedback in brain state was also shown in a model of rodent active whisking, where brief interruptions in this feedback enhanced signal-to-noise ratio for detecting touch. Finally, in monkey visual fixation, artificial CLE feedback suppressed stimulus-specific neuronal activity and improved behavioral performance. Our findings show that the environment mediates continuous closed-loop feedback that controls neuronal gain, regulating brain state, and that brain function is an emergent property of brain-environment interactions.
q-bio.NC
q-bio
Brain State Control by Closed-Loop Environmental Feedback Christopher L. Buckley1,2,*, Satohiro Tajima1, Toru Yanagawa1, Kana Takakura1, Yasuo Nagasaka1, Naotaka Fujii1, and Taro Toyoizumi1,* 1 RIKEN Brain Science Institute, Japan 2 University of Sussex, UK Summary Brain state regulates sensory processing and motor control for adaptive behavior. Internal mechanisms of brain state control are well studied, but the role of external modulation from the environment is not well understood. Here, we examined the role of closed-loop environmental (CLE) feedback, in comparison to open-loop sensory input, on brain state and behavior in diverse vertebrate systems. In fictively swimming zebrafish, CLE feedback for optomotor stability controlled brain state by reducing coherent neuronal activity. The role of CLE feedback in brain state was also shown in a model of rodent active whisking, where brief interruptions in this feedback enhanced signal-to-noise ratio for detecting touch. Finally, in monkey visual fixation, artificial CLE feedback suppressed stimulus-specific neuronal activity and improved behavioral performance. Our findings show that the environment mediates continuous closed-loop feedback that controls neuronal gain, regulating brain state, and that brain function is an emergent property of brain-environment interactions. Introduction The repertoire of animal behavior involves both passive and active interactions of the brain with the environment. Passive interactions are driven by the environment and convey signals for passive sensing and alert. Active interactions, in contrast, are bidirectional, mediating movement and goal-directed behaviors. During active interactions sensory input from the environment is shaped by motor actions, and sensory percepts inform future motor commands, forming a closed-loop between action and perception. This closed-loop environmental (CLE) feedback is central to the production of motor control behaviors (Wolpert and Ghahramani, 2000; Kawato, 1999; Scott, 2004) and active sensing behaviors 1 such as saccading and sniffing (Smith 1962; Von Holst, 1954; Crapse and Sommer, 2008). Yet, how this CLE feedback impacts on brain state, i.e., brain-wide neuronal dynamics and processing, has received relatively little attention. Brain state, typically characterized by the degree of synchronously fluctuating neuronal activity as reported by electroencephalography (EEG), local field potentials (LFP), electrocortiography (ECoG), membrane potential, and population spiking activity, is strongly modulated by behavioral context (Buzsáki et al., 2012; Harris and Thiele, 2011). One of the earliest experimental examples of this phenomenon was the demonstration that alpha frequency power (7.5–12.5Hz) in EEG recordings is enhanced when subjects close their eyes or during periods of drowsiness (Berger, 1929). More generally recent experiments in rodents have demonstrated that synchronous low-frequency fluctuations (a synchronized state) are typically suppressed (moving to a desynchronized state) when animals transition from quiet attentive behaviors, where animals are largely passive, to active behaviors such as running or whisking (Crochet and Petersen, 2006; Niel and Stryker, 2010; Schneider et al., 2014; Otazu et al., 2009; Zagha et al., 2013). Brain states not only alter spontaneous brain dynamics, but also sensory representations, presumably allowing brain function to adapt to behavioral context. For example, it has been shown that the onset of running amplifies cortical responses to visual and auditory stimuli (Niell and Stryker, 2010, Fu et al., 2014; Schneider et al., 2014; McGinley and McCormick, 2014). In addition, cortical responses to punctuate stimuli are larger during synchronized states (Poulet and Petersen, 2008; Fanselow and Nicolelis, 1999; Castro-Alamancos, 2004, Otazu et al., 2009). These brain state changes are typically thought to happen across the whole brain (Berger, 1929, Steriade, 2001; Fu et al., 2014; McGinley and McCormick, 2014) and are cross-modal (Otazu et al., 2009). 2 Several internal mechanisms have been implicated in these changes including neuromodulatory pathways (Goard and Dan, 2009; Polack et al., 2012; Pinto et al., 2013; Fu et al., 2014), thalamo-cortical projections (Poulet et al., 2012), and intracortical feedback (Zagha et al., 2014; Schneider et al., 2014). Changes in the external sensory environment also play an important role. Under passive sensing conditions sensory stimulation can shift neural activity in the visual cortex between asynchronous and synchronous regimes (Tan et al., 2014) and more generally sensory stimuli quench neural variability (Churchland et al., 2010). More relevant to the present study, brain state transitions often coincide with the onset of active behaviors, which are characterized by the presence of reafferent input (sensory input resulting from one's own action, mediated by the environment). It is well known that reafferent input strongly influences neural activity (Curtis and Kleinfeld, 2009; Urbain et al., 2015; Brooks et al., 2015) and modulate sensory responses (Diamond et al., 2008). However, the role played by reafferent input in regulating brain state is not well understood. In this paper, we combine theoretical and experimental approaches to dissect the effect CLE feedback mediated by reafferent input, as oppose to sensory input per se, has on modulating brain state. In doing so we elucidate a novel function that CLE feedback plays in sensing and behavior. Results During active behavior, sensory input to the brain is directly shaped by motor actions, and reciprocally, motor actions are informed by prior sensations forming a closed-loop between the brain and the environment. Under such conditions, it is possible to distinguish two sources of sensory input (Von Holst, 1954), see Figure 1A. First, exafferent input describes sensory signals that originate in the environment, but which are completely external to the brain (Von Holst, 1954). Second, reafferent input describes sensory signals, which, while mediated by the environment, are a direct consequence of an animal's own actions and constitute a sensory feedback signal to the brain (Von Holst, 1954). While it is known that 3 sensory input in general can influence brain state variability (Churchland et al., 2010), it is not clear whether exafferent and reafferent input play distinct roles in determining these changes, or whether only total synaptic input is of importance. To examine the effect of each type of input on brain state, we distinguish three types of brain-environment interaction. First, an open-loop condition, where sensory input is not coupled to motor output and thus the brain receives sensory input independent of its own activity (Figure 1B). In this condition, reafferent input is absent and the brain receives only exafferent input. Second, a closed-loop condition, where the brain interacts with the environment in a closed-loop thus coupling motor action and sensory perception (Figure 1C). In this condition, the brain receives reafferent input in addition to any possible exafferent input. Lastly, we define a replay condition, where again the brain operates in an open-loop but where reafferent input experienced during a prior closed-loop condition is recorded and replayed as exafferent input (Figure 1D). In this condition, the total sensory input is identical to that in the closed-loop condition, but reafferent input is absent. Specifically, the sensory input during the replay condition is not a consequence of motor actions within that condition. Therefore, this condition serves as a strong control allowing us to distinguish precisely the contribution made by reafferent input rather than that of total sensory input, to the brain state. To understand the possible effects that CLE feedback could have on brain activity we introduce a simple idealized model. We start by describing collective neuronal activity, e.g., an EEG signal, of an open-loop brain (denoted as 𝐵𝑜(𝑡)), in terms of a first-order differential equation, 𝑑𝐵𝑜(𝑡) 𝑑𝑡 = − 𝐵𝑜(𝑡) 𝜏 + 𝜉𝑜(𝑡), where 𝜉𝑜 is white noise of instantaneous variance 𝜎2 generated inside the brain, 𝑡 is time, and 𝜏 is the time constant of the system. The autocorrelation peak (instantaneous variance) of variable 𝐵𝑜 is given by Peak𝑜 = 𝜎2𝜏/2 (see, Figure 1B for open-loop time trace). To 4 describe neuronal activity in the closed-loop condition (denoted as 𝐵𝑐(𝑡)), we approximate reafferent input as a linear function of the brain variable, i.e., 𝑑𝐵𝑐(𝑡) 𝑑𝑡 = − 𝐵𝑐(𝑡) 𝜏 + 𝑤𝐵𝑐(𝑡) + 𝜉𝑐(𝑡), where 𝑤 scales the strength of reafferent input and 𝜉𝑐 is again white noise of instantaneous variance 𝜎2. (Later we also explore a more realistic reafferent input that involves filtering and delay). In this condition, the continuous cycles of reafferent input constitutes a CLE feedback signal to the brain. The presence of this CLE feedback changes the effective time constant to 𝜏eff = 𝜏/(1 − 𝑤𝜏) or, equivalently, it the changes the system's gain. We can quantify this change by using an expression for the autocorrelation peak of the system, as follows, Peak𝑐 = Peak𝑜/(1 − 𝑤𝜏). In particular, we find that CLE feedback suppresses fluctuations within the brain if this feedback is negative (𝑤 < 0; see Figure 1C for closed-loop time trace). In contrast, we can describe neuronal activity in the replay condition (denoted as 𝐵𝑟(𝑡)) as: 𝑑𝐵𝑟(𝑡) 𝑑𝑡 = − 𝐵𝑟(𝑡) 𝜏 + 𝑤𝐵𝑐(𝑡) + 𝜉𝑟(𝑡), where 𝜉𝑟 is again white noise of instantaneous variance 𝜎2. Here, the brain receives the same total sensory input as in the closed-loop condition, 𝑤𝐵𝑐(𝑡), but as exafferent input, i.e., it depends on the dynamics of 𝐵𝑐(𝑡), and not on 𝐵𝑟(𝑡). Again we can calculate the autocorrelation peak of the brain variable in this condition as Peak𝑟 = Peak𝑐 + Peak𝑜 ∗ 2𝑤𝜏/(𝑤𝜏 − 2). Notably, even though the brain receives exactly the same total sensory input in the replay and closed-loop conditions the amplitudes of fluctuations are not the same. In particular, if CLE feedback is negative, we obtain peak𝑐 < peak𝑜 < peak𝑟 (see, Figure 1D for replay time trace). In summary, this simple model suggests that CLE feedback constituted by continuous reafferent input could have a profound effect on neuronal fluctuations and thus 5 could be implicated in the modulation of brain state. In the next section, we examine this hypothesis using experimental data. Negative CLE feedback suppresses synchronous neural fluctuations We tested our predictions by analysing two-photon imaging data recorded from larval zebrafish behaving in a virtual flow simulator (Ahrens et al., 2012), a setup that allowed us to quantify the differences between the closed-loop and replay conditions directly (Figure 1). In this setup, bouts of activity recorded from motor nerves along the spine of a paralyzed fish were translated into a backward drift of a visual grating, simulating forward swimming. This allows the fish to maintain its perceived horizontal position by swimming against water flow (Ahrens et al., 2012), (Figure 2A). In a transgenic fish expressing the calcium indicator GCaMP2 brain-wide calcium activity was monitored using a two-photon microscope to scan single planes in the brain. In a closed-loop condition, the fish actively maintain their position in the virtual environment. In a replay condition, the same fish receives a replay of the closed-loop visual stimulus without real-time visual feedback (Figure 2A), (Ahrens et al., 2012). We found that neural dynamics between the two conditions were significantly different despite the fact that both the closed-loop and replay conditions involve identical sensory input (Figure 2B). Notably, the only information the fish received about oncoming flow was visual, i.e., there was no proprioceptive input as the fish was paralyzed (Ahrens et al., 2012). Individual neurons were heterogeneous across the whole brain, but on average low frequency (0.01 - 0.15 Hz) fluctuations were suppressed (p = 0.046, sign test) and neurons were decorrelated (p = 0.005, sign test) under the closed-loop condition compared to the replay condition (Figure 2B). Changes in the geometric mean of low frequency fluctuations and correlation were highly correlated in each pair of cells (r=0.69, p<10-10, Spearman's rank correlation), suggesting a common cause (see Figure 2C). The decorrelation effect was not an artifact of measurement noise, which may dominate at high frequency because the result was robust to the removal of low-level calcium activity by thresholding (see Supplementary Figure S1A). This change aligns with brain state transitions at the onset of active behavior that has been reported in other species (see Poulet et al., 2012 or Harris and Thiele, 2011 for a review). This 6 suppression of low frequency power and correlation could be caused by an efference copy signal (signal encoding intended motor action) that suppresses synchronous neuronal fluctuations (Zagha et al., 2014; Schneider et al., 2014). However, while motor activity levels were higher in the closed-loop condition than the replay condition, this did not explain the difference in neuronal fluctuations. Specifically, the motor activity level positively correlated with the changes in low frequency power (Supplementary Figure S1B, top panel; r = 0.18, p < 10-2, Spearman's rank correlation) and was not significantly correlated with changes in pairwise correlations between cells (Supplementary Figure S1B, bottom panel r = 0.03, p > 0.5, Spearman's rank correlation). The only difference between the closed-loop and replay conditions was the presence of CLE feedback; thus it can be assumed that this must play a causal role in these changes. To investigate this further, we quantified CLE feedback by first estimating how fish behavior affects neural activity, and conversely, how neural activity drives their behavior. To this end we calculated linear filters that characterize dynamic interactions between single neuron activity (the brain variable: B) and swimming power as quantified by the activity of motor neurons (a putative environmental variable: E) (see Methods). Figure 3A (inset) schematically shows their interaction under each condition. In the closed-loop condition B and E interact mutually; this has the potential to confound calculation of independent causal filters. Consequently, we calculated filters based on the replay condition (however, see below for filters computed based on the closed-loop condition). In this condition, the brain variable B' is driven by the recorded CLE variable E, which in turn generates its own environmental variable E', without involving CLE feedback. Specifically, for each observed neuron, we computed a linear filter that predicts B' from E and a linear filter that predicts E' from B'. While these filters are neuron-dependent, the average normalized filters showed clear net effects (Figure 3A). On average, across cells, the interaction from the brain to the environment, B'→E' (Figure 3A, purple solid line), was strongly positive, indicating that an 7 increase in neural activity positively drove motor behavior. However, we found that interaction from the environment to the brain, E→B' (Figure 3A, pink solid line), was net negative, indicating that an increase in motor behavior on average suppressed neural activity. It is reasonable to assume that the same filters (circuits) operate in the closed-loop condition. If this is the case, CLE feedback for each cell in the closed-loop condition is characterized by the convolution of the two filters, i.e., by the convolution of the B'→E' and E→B' interactions computed for each cell. The convolution of these filters was on average also negative (Figure 3A, cyan dashed line), peaking at about 1 s. This suggests, that on average, increases in neural activity self suppressed after 1 s due to negative CLE feedback. Notably, the negative CLE feedback interactions were also confirmed by behavioral analysis, in which the the E→E' filter was estimated directly (see Supplementary Figure S1C), which supported the robustness of our results. Next, we considered what the consequence of this negative CLE feedback on each cell's activity would be. Our conceptual model, Figure 1, suggests that fluctuations in neural activity should decrease if a cell receives negative CLE feedback. In this case the estimated CLE feedback in each cell should predict to what extent fluctuations in low frequency power should change in the closed-loop condition relative to the replay condition (see Methods). As theoretically expected, Figure 3B demonstrates that the predicted degree to which a neuron's activity was suppressed during the closed-loop condition relative to the replay condition was highly correlated with what was actually observed (r = 0.39, p < 10-8, Spearman's rank correlation). Note that neural activity in the closed-loop condition was not used to fit each filter. Although the majority of cells across the fish brain were suppressed by the behavior, the top 10 percentile of cells that were both strongly suppressed, and strongly involved in, negative CLE feedback were clustered in the cerebellum (Figure 3C), a brain area implicated in sensory-motor planning and coordination (Scott, 2004). This supports the idea that the 8 cerebellum plays a central role mediating negative closed-loop interaction between the brain and the environment by converting sensation into action in fish during optic flow stabilization. We then investigated whether the dynamic relation between neuronal activity and motor activity were also affected by CLE feedback. To do this, we quantified the brain/environment dynamics for each neuron by naively evaluating the E→B filter in the closed-loop condition, and compared this to the corresponding E→B' filter in the replay condition (Figure 3D). These two filters were generally distinct in the observed neurons, but were particularly so for those cells that were strongly stabilized by CLE feedback (Figure 3D; Inset). In these neurons, while the E→B' filter from the replay condition was approximately causal, as expected, the closed-loop filter E→B had an additional acausal component. We suggest that this latter filter reflects the closed-loop interaction between the brain B and the environment E. To test if the closed-loop interaction could explain this discrepancy, we theoretically predicted the E→B filter in the closed-loop condition based on data from the replay condition, i.e., using both the E→B' and B'→E' filters (see Methods). We found that, on average, this closed-loop effect could account for the difference between E→B and E→B' as calculated for the closed-loop stabilized cells (Figure 3D; Inset). To quantify this for each cell, we calculated the mean square error between the predicted and E→B' filters and between the E→B and E→B' filter. The ratio of these two error terms then quantifies the fraction of the mean square error that was explained by the prediction. This ratio was significantly less than one (median = 0.80, p < 10-11, sign test), indicating that the prediction was more accurate when the closed-loop effect was included, and the ratio was positively correlated (r=0.25, p<10-13, Spearman's rank correlation) with the degree to which individual cells were stabilized in the closed-loop 9 condition (Figure 3D). Altogether, these results indicated that neuronal dynamics, as well as its relation to sensory stimulus and behavior, not only depend on brain circuits, but are also dynamically shaped by the mutual interaction between the brain and the environment. CLE feedback explains the difference between active and passive sensing Behaviors not only induce brain state transitions, but also differentiate active sensing from passive sensing. To investigate whether these differences could be accounted for by CLE feedback, we examined a well-studied model of a brain state transition in the rodent whisker system. Specifically, we consider CLE feedback to the rodent brain mediated by whisking vibrissa. Although more precisely we should refer to this as closed-loop body/environmental feedback, we retain CLE acronym by generalizing the notion of environment to refer to all processes outside of the brain highlighting the generality of our theory. Whisking not only changes brain state (Crochet and Petersen, 2006; Poulet and Petersen, 2008), but also reduces the sensitivity of the cortex to passive whisker deflections (Crochet and Petersen, 2006; Crochet et al., 2011). However, interestingly, robust responses are maintained for active contact events when the whisker collides with an object placed in the whisk field (Crochet and Petersen, 2006; Crochet et al., 2011). Furthermore, in an active touch condition where the whisker repeatedly collides with an object both intra-neuronal correlations and low frequency power of membrane potential recovers close to their passive state values (Crochet et al., 2011). We theoretically investigated whether these phenomena could be explained by the hypothesis that vibrissa dynamics mediates negative CLE feedback to the brain, influencing cortical dynamics in a way analogous to fish swimming behavior. Notably, while multiple different mechanisms are involved in brain state transition (Goard and Dan, 2009; Polack et al., 2012; Pinto et al., 2013; Fu et al., 2014; Poulet et al., 2012; Zagha et al., 2014; Schneider et al., 2014), for which sensory input is not always necessary (Poulet and Petersen, 2008; Poulet et al., 2012), this does not exclude the role of CLE feedback under physiological conditions. Indeed, although brain state transitions can happen at whisking onset even after the sensory nerve is cut (Poulet and Petersen, 2008), our analyses revealed that the latency of brain state transition, measured by whisking-related reduction in low frequency power of 10 cortical LFP or in thalamic spiking rate, was delayed under this condition as compared to the control condition (Supplementary Section S2). This supports a physiological role for sensory input in the brain state transition. Here, we construct a simple model of the vibrissa system to investigate the possible role of CLE interaction. The model comprises excitatory and inhibitory cortical populations dynamically interacting with a single vibrissa (Figure 4A; see Methods for details). In our model, synchronous membrane potential fluctuations arise endogenously from interplay between the build up of excitatory cortical activity by recurrent connections and their eventual suppression by adaptation after ca. 1 s in each cell (see Methods). To model the brain/body interaction, we assumed that cortical excitatory neurons drive whisker motor behavior and both excitatory and inhibitory neurons receive the sensory input that reflects whisker angle. Based on our theory, Figure 1, we hypothesized that synchronous neural fluctuations could be reduced during whisking by negative CLE feedback. We modeled this by assuming that cortical excitatory neurons are activated by whisker retraction but this activation drives whisker protraction (see Figure 4A). (Note the specific biological implementation of the negative CLE feedback is not central to our claims, and there are several other plausible schemes, see Discussion and Supplementary Section 6.) We also model a central pattern generator (CPG, likely located in the brain stem) (Hill et. al, 2011) that, once turned on, rapidly cycles the position of the vibrissa back and forth at around 10Hz. Sensory input during contact events involves both contact-related and whisker angle-related signals (Diamond et al., 2008). To capture this sensory input we modeled a single whisker as two stiff mass-less sections connected by hinges at the center and the base, which are constrained by muscles (simple torsion springs). Whisking is implemented by driving the equilibrium position of the base spring. The center spring has an equilibrium value of zero angular displacement and thus tends to align both sections (see Supplemental Information Section 3). We start with a simple non-bending stiff whisker but later vary the stiffness 11 parameter (see Methods). A horizontal solid wall is placed above the whisker and, as the whisker collides with the wall, the whisker tip stops, see Figure 4A, interrupting CLE feedback about whisker angle. Whisker contact not only involves a modification of CLE feedback but also invokes exafferent input, or contact-detection signal, that results from the stereotypical response of pressure sensitive cells in the trigeminal ganglion (Szwed et al., 2003). We model this as a brief square wave pulses (ca. 25 ms) triggered by each contact event (see Methods). Our key hypothesis, i.e., that whisker position mediates negative CLE feedback to cortical neurons, allowed us to reproduce rodent brain state findings that are observed under three behavioral conditions (the quiet condition, the whisking condition, and the active touch condition; Figure 4B [Crochet et al., 2011]). Specifically, the synchronous low frequency membrane potential fluctuations generated during the quiet condition (open-loop: the whisker position is fixed at 0) were significantly suppressed during the whisking condition (closed- loop: the whisker is driven by the CPG and cortex) (Figure 4C). In effect, the negative CLE feedback reduced the gain of the cortical system, see conceptual model Figure 1, and replaced prominent (ca. 1 Hz) synchronous fluctuations of the membrane potential with fast (ca. 10 Hz), but weak, fluctuations locked into the whisking cycle. As a result, the average inter-neuronal correlation of membrane potential for pairs of neurons was also suppressed (Figure 4D). During the active touch condition, occasional whisker touch events stopped the whisker tip on the object's surface (Figure 4B; touch events marked by red bars). Both the low frequency power and cross-correlation of membrane potential fluctuations were partially recovered during the active touch condition in agreement with experimental results (Crochet et al., 2011) (Figure 4C and D, red lines). We hypothesized that this recovery was a consequence of intermittent interruption of negative CLE feedback during whisker contact events. To verify this, we examined whether this result depended on the stiffness of the whisker. We also 12 hypothesized that, if the whisker is very flexible, then the base angle would change continuously, despite contact of the tip, fully preserving CLE feedback. In contrast, a very stiff whisker would come to complete rest on the wall during each contact maximally interrupting CLE feedback (see also Supplementary Figure S3B). We found that the recovery of 1-5 Hz power was indeed predominantly caused by the interruption of CLE feedback and not touch- evoked exafferent input (see Supplementary Figure S3C) demonstrating that these transient interruptions of CLE interaction can rapidly switch the active brain state to a passive brain state. This model also accounted for observed brain state dependent changes in sensory processing (Crochet et al., 2011) without assuming additional mechanisms (Lee et al., 2008; Nguyen and Kleinfeld, 2005). In agreement with experimental results we found that passive exafferent input evoked large sensory responses in the quiet condition but markedly smaller responses during the whisking condition (Figure 4E) (Crochet et al., 2011). Again, this was because negative CLE feedback decreases the gain of the cortical circuit (Supplemental Section 4). However, in agreement with experimental data, cortical neurons exhibited more reliable responses to active touch events (Figure 4B and E). This was because the negative CLE feedback, that suppressed neural fluctuations during whisking, was transiently removed during active touch events allowing endogenous brain dynamics to amplify the cortical response to exafferent input (Figure 4E and F). Consequently, active touch events combined large sensory evoked responses (signal) and low background fluctuations (noise). To quantify this effect, we computed a discriminability index (see Methods) that measures the separation between the distributions of membrane potentials in the presence or absence of sensory events. The value of the index was similar for perturbations in both quiet and whisking conditions, i.e., although the deflection-evoked response (signal) was greater in the quiet condition, so were background fluctuations of membrane potential (noise) (see Figure 4E and F). In contrast, the discriminability index was greater for active touch events (Figure 13 4E and F). Hence, the model suggests that cortical neurons are selectively sensitive to interruption of the animal's own active sensing. All these results are intuitively reproduced in a simplified model of CLE interaction, demonstrating the robustness of our findings (Supplemental Section 4). Neurofeedback modulates brain dynamics and behavior We further tested a core prediction of our theory by examining the impact of an artificially constructed CLE feedback on brain dynamics. Specifically, we investigated whether we could use an ECoG-based fast neurofeedback technique to modulate brain dynamics and subsequently behavioral performance in primates. This setup allowed us to observe the impact of CLE interaction without involving efference copy signal (which can induce brain state transition; Zagha et al., 2014; Schneider et al., 2014) and to investigate the possible clinical relevance of CLE feedback. We implemented real-time neurofeedback between the visual areas (occipitotemporal and parietal cortices) of a fixating macaque monkey (equipped with a 128-channel ECoG on the cortical surface, Supplementary Figure S5A) and a visual stimulus (equivalent to the environment in the fish study or vibrissa feedback), while the monkey was fixating (Figure 5A). During the presentation of a visual grating stimulus (Figure 5A), we decoded orientation (vertical or horizontal grating), in real-time, from ECoG activity with a support-vector machine. During neurofeedback sessions, a computer monitored the output of the classifier (the decision value), a nonlinear projection of the ECoG activity, and then modified the stimulus presented on the screen in real-time. A large and positive (or negative) decision value indicated that ECoG activity was likely driven by the horizontal (or vertical) grating stimulus, respectively. When the decision value exceeded a positive threshold, the computer showed a vertical grating for the subsequent 100 ms period; vice versa, exceeding a negative threshold triggered the presentation of a horizontal grating (see Methods). If the decision value fell between these two thresholds, a grey screen was presented. This protocol effectively approximates negative feedback (the closed-loop condition) on the decision value dynamics of the animal (however see Discussion). To 14 distinguish between the influence of neurofeedback from the visual stimuli alone, we used a control condition in which the monkey was presented with a recording of the visual stimuli that emerged during the closed-loop condition (a replay condition). Thus, as in our conceptual mode (Figure 1A), and analogous to the fish experiment (see Figure 2A), corresponding closed-loop and replay conditions share identical visual stimuli and only differed by the presence of neurofeedback. Again, this allowed us to examine the exact effect of CLE feedback, rather than sensory input on brain activity and behavior. We observed that neurofeedback immediately altered the dynamics of the decision value and thus brain dynamics, without requiring any prior training or habituation of animals. The power spectrum of the decision value below 5 Hz significantly differed under the two conditions (Figure 5B). While the 0-1 and 3-4 Hz (summarized in Figure 5C) frequency ranges were suppressed, the 1-3 Hz frequency range was enhanced during the closed-loop condition relative to the replay condition. This oscillatory modulation of the power spectrum could have arisen from delay (ca. 100 ms) in negative CLE feedback, but it would also depend on various factors, including the kinetics of the decision value, and early visual stream processing. However, the exact mechanisms underlying the changes in the power spectrum are not the focus of this study. We found that not only were the dynamics of the decision value altered by neurofeedback, but also that there was significant improvement in the monkey's fixation performance (Figure 5D). Specifically, the distance of the eye position from the fixation point was quantified and compared under the two conditions. The fixation performance at 1600-2100 ms after the appearance of the fixation point (about 300-900 ms after typical grating onset; Supplementary Figure S5C) in the closed-loop condition was enhanced (Figure 5E). To establish a quantitative relationship between changes in neural dynamics and behavior, we analyzed the trial-to-trial correlation between the power spectrum of the decision value and fixation performance. The reduction of fluctuation in the decision value at specific frequencies correlated with improved fixation performance (Figure 5F). In particular, the neurofeedback-induced reduction of decision value fluctuations in the 3-4 Hz 15 frequency range (estimated from data before the behavioral improvement; see Supplementary Figure S5D) showed a significant correlation with improved fixation performance (Figure 5G). Taken together, these results indicate that in primates artificial neurofeedback can change brain dynamics related to sensing and task-related behavior, which was consistent with our fish and rodent studies. Although the causal chain of events underlying these changes may also involve internal brain mechanisms for state transitions (e.g., neuromodulators; see Discussion), we assert that neurofeedback constitutes the primary cause of these changes as it is the only difference between each condition. Discussion Brain state transitions triggered by the onset of active interactions between the brain and environment, represent a major neuronal mechanism shaping sensing and behavior. In this study, we use both theory and experiment to support the idea that negative CLE feedback inhibits network gain, which in turn, suppresses synchronous neuronal fluctuations and sharpens sensory responses. We generalize and support the theoretical framework in three diverse animal model systems summarized in Figure 6. In each system, we show that negative CLE feedback regulated real-time brain state and animal behavior. Specifically, CLE feedback quantitatively predicted cell-specific suppression of neural oscillations in zebrafish, enhanced signal-to-noise ratio for active sensing in rodent, and enhanced task performance in primate vision. The importance of using naturalistic sensory stimuli to study and manipulate brain state dynamics is widely demonstrated (Felsen and Dan, 2005). However, an important prediction of our theory (Figure 1, Supplemental Information Section 7), supported by our experimental findings (Figures 2, 3, and 5), is that brain dynamics during active sensing cannot be fully recapitulated or re-encoded, even if the same sensory input is precisely recorded and replayed back into a passive brain. These results provide evidence that brain state neurometrics and behavioral psychometrics during active behaviors can only be accurately understood by a quantitative account of ongoing brain-environment interactions (O'Regan 16 and Noë, 2001). Neuronal gain control by CLE feedback The formal component of our theory, i.e., that CLE feedback can modulate a system's gain, is well documented in dynamical systems theory and control theory (e.g., Aströmand Murray, 2010). This gain control occurs even though the instantaneous effect of the pathways mediating feedback is purely additive (c.f. Figure 1) because the effect of repeated cycles of feedback accumulates over time. For example, input 𝐼 to a linear dynamical system with feedback strength 𝑏 makes a first direct contribution 𝐼 to its response, but then also makes subsequent contributions as this initial response cycles around a feedback loop-a contribution of 𝑏𝐼 after one cycle, 𝑏2𝐼 after two cycles and so on. The cumulative sum of these contributions I + 𝑏𝐼 + 𝑏2𝐼 +. . . = 𝐼/(1 − 𝑏) is equivalent to divisively scaling the input magnitude by a factor that depends on the feedback strength, i.e., effectively changing the system's gain. Thus, a constitutively active closed-loop feedback that mediates multiple action-perception cycles is essential for the form of gain control we propose. This means that discrete and intermittent involvement of reafferent input does not imply gain modulation. For example, the classical reafference principle explains neuronal responses by a one-time detection of the mismatch between an efference copy (predicted) and reafferent (actual) (Von Holst, 1954). However, this situation is likely an inaccurate idealization to describe the closed-loop systems studied here. For example, in the zebrafish system, swim bouts typically occur every 700 ms and this interval closely overlapped with the peak of the estimated environmental feedback interaction (Figure 3A and Supplementary Figure S1C). Hence, the neural responses in the fish experiment suggest a more dynamic system, where neural activity evoked by many cycles of action and sensation are continuously and mutually interacting. Neuronal mechanisms of negative CLE feedback The presence of continuous negative CLE feedback during active behavior is fundamental for our theory. Although, as a higher order theory, it is agnostic to the detail of the neural implementation, we 17 discuss below the implications for each system. In zebrafish, the presence of negative CLE feedback during fish swimming behavior would seem a priori necessary for optic-flow stabilization behavior because the fish must act in opposition to perceived optic flow in order to minimize horizontal displacement (Wolf et al. 1992, Fry et al., 2009). Interestingly, neurons that received strong negative CLE feedback and were substantially stabilized were located in the cerebellum (Figure 3C) consistent with a theoretical viewpoint that the cerebellum mediates this optomotor response by converting sensation into action; e.g., by completing the action-perception cycle (Ito 1972, Lisberger et al. 1987, Kawato 1999). Our rodent whisker model explores the overall effects of negative CLE feedback mediated by the cortical-whisker circuit during active whisking. As in the fish study, the theoretical assumption of negative CLE feedback is consistent with the idea that the barrel cortex is involved in control. Specifically, the barrel cortex comprises a nested set of servo control loops that regulate various aspects of whisker dynamics (Ahissar and Kleinfeld, 2003). At the level of the whole vibrissa system, multiple parallel and nested feedback loops, both positive and negative (Ahissar and Kleinfeld, 2003) most likely exist. While our model is abstract in terms of the known complexities of vibrissa system anatomy and exact concordance of the model with vibrissa system anatomy is beyond the scope of this paper, we provide several possible schemes for experimentalists to examine in Supplemental Information Section 6. The artificial CLE feedback in the primate experiment implements negative feedback, although this is less obvious than in our other studies and simple models. A system is defined to be under negative feedback if this feedback causes perturbations to the system's state to decay faster in time and suppresses fluctuations due to noise. Hence, neurofeedback in the monkey experiment is a hybrid of positive and negative feedback-the feedback is negative in the 0-1 Hz and 3-4 Hz frequency bands and positive in the 1-3 Hz frequency band, because the feedback reduces or increases fluctuations of the decision value in these frequency bands, respectively (Figure 5B). In this experiment, the system's state was quantified by the decoder's decision value, with positive and negative values indicating 18 vertical-like and horizontal-like brain activity patterns, respectively. Due to experimental constraints of delays in neurofeedback processes and unidentified dynamic properties of the brain, imposing purely negative neurofeedback in this experiment was not possible. However, negative CLE feedback in the 3-4 Hz frequency band is strongly correlated with enhanced fixation performance (Figure 5F), supporting its behavioral significance. Comparison to internal mechanisms of brain state transitions The proposed mechanism of brain state transitions by CLE feedback is qualitatively different from the many other mechanisms governing brain state that operate inside the brain (Goard and Dan, 2009; Polack et al. 2012; Pinto et al., 2013; Poulet et al., 2012; Zagha et al., 2014). Specifically, the CLE feedback mechanism involves the dynamic coordination between brain activity and body/environmental dynamics (c.f. Figure 3D), a continuous reciprocal interaction that is critical for many forms of goal-directed behavior. In this regard, our theory directly links a mechanism of brain state transitions to behavior and is closed from an explanatory view, i.e., without assumptions about upstream causes. In our zebrafish study, an alternative interpretation of our findings is that correlated neural fluctuations are suppressed directly by an efference copy signal (Zagha et al., 2014; Schneider et al. 2014) due to greater motor commands in the closed-loop condition (Supplementary Figure S2). However, in opposition to this view, we found that motor activity correlated positively with fluctuations in each cell, and were not associated with pairwise neuronal correlations (Supplementary Figure S2). Hence, we cannot account for the results by any known functions of efference copy signals. Moreover, the primate neurofeedback allowed us to examine the impact of CLE feedback in the absence of the overt involvement of the motor system, i.e., feedback was only dependent on readout from visual areas. This allows us to conclude that CLE feedback alone, in the absence of noticeable motor activity, is sufficient to modulate brain activity and behavior. 19 Importantly, brain state control by CLE feedback is not mutually exclusive with other mechanisms, such as thalamo-cortical input (Poulet et al., 2012) or neuromodulation (Goard and Dan, 2009; Polack et al. 2012; Pinto et al., 2014; Fu et al. 2014) because these may also be involved in mediating the action-perception cycle. Furthermore, brain state transitions also occur in the absence of CLE feedback, such as open-loop behaviors (e.g., the onset of running that does not change the visual screen or grooming) (Niell and Stryker, 2010, Fee et al. 1997), during sleep (Vyazovskiy et. al 2011; Steriade, 2005), or under dissection of the sensory nerve (Poulet and Petersen, 2008; Poulet et al., 2012). Mechanisms underlying brain state transitions are likely to be highly redundant even in the absence of essential mechanisms, such as thalamo-cortical input (Poulet et al., 2012) or corollary discharge (Schneider et al. 2014), albeit involving further delay (see also Supplementary Section 2). Such functional redundancy may help to maintain the stability of brain state (Fu et al. 2014; McGinley et. al, 2015; Otazu et al. 2009; although with exceptions, see Vyazovskiy et. al 2011). Furthermore, the relative importance of internal and external mechanisms might adaptively change in an experience- dependent manner (Nachestedt et. al 2014). Reafferent feedback in active sensing and revision of the reafference principle In the whisking model, we proposed that the regulation of cortical gain by CLE feedback could explain enhanced active touch. Specifically, negative CLE feedback during whisking reproduces suppressed fluctuations and reduces responses to passive whisker stimulation (see Figure 4F). However, robust neuronal response to active touch events could be explained by the interruption of CLE feedback when the whisker is driven into an external object. Such events interrupt CLE feedback, transiently releasing the cortex from a low gain state and enhancing sensory responses to salient sensory stimuli (Figure 4E, Supplemental Information Section 4). This mechanism for active touch contrasts with the account of sensory processing suggested by the reafference principle (Von Holst, 1954), which postulates that motor efference is discounted from sensory input, allowing animals to sense exafferent signals (externally caused sensory input) without being confounded by the consequences of their own motor actions. In contrast, our theory suggests that the sensory system is insensitive to pure exafference during active sensing (Figure 4F), but sensitive to the interruption of 20 reafference which may allow animals to focus attention on the consequences of their own motor actions. While it is straightforward to generalize this idea to other tactile systems, its implication for other sensory modalities is less clear. However, in theory, CLE feedback could be interrupted anywhere along the action-perception cycle, thus dynamically regulating neuronal gain. The timely interruption of this feedback could serve as a general mechanism for temporarily accentuating neuronal responses against a background of reduced noise (Scott, 2004, Hafed et al., 2011). For example, cerebellum neurons, which are strongly involved in the sensory-motor cycle, could be suppressed in anticipation of salient sensory events by a relevant brain area, such as the reticular formation (Kinomura et al. 1996). Rapid neurofeedback and virtual reality-based behavioral enhancement The primate findings demonstrate that neurofeedback can modulate brain dynamics (Figure 5B and C) and enhance task-related behavior (Figure 5D, E, F, and G). These results portend the use of CLE feedback as an interventional tool for behavioral enhancement in closed-loop feedback therapies. Unlike more conventional neurofeedback techniques (Zoefel et al., 2011, Shibata et al., 2011) that require training periods to affect behavior, our technique enhanced fixation performance within seconds and required little supervision. This rapid neurofeedback was achieved by crucial differences in the design between this technique and conventional neurofeedback. Conventional protocols require conscious/unconscious human learning that is supported by an appropriate neurofeedback signal. In contrast, our protocol directly intervened on the fast time scale of the ongoing sensory stream. Our approach has the potential to be more stable, with more rapid, user-independent, and effective behavioral control than current conventional decoded neurofeedback methods. The neurophysiological reason underlying the improved fixation performance is not clear. However, the high-contrast grating stimulus likely serves as a distractor for the monkey and makes fixation more 21 difficult. During the neurofeedback protocol, the 3-4 Hz frequency component of this distractor signal was suppressed by the feedback, possibly inhibiting the formation of a grating perception. If so, it is possible that such CLE feedback may help to allocate attention in physiological settings. Specifically, our model predicts that, animals may show less attention to sensory modalities that provide ongoing negative CLE feedback. A full investigation of these issues will require further work. These findings emphasize the importance of the development of natural and virtual reality systems that provide well-controlled quantitative measurements of CLE interactions. Such novel systems would be useful for studying how healthy and disease brain physiology emerges from real-time brain-body- environment interactions and may suggest new methods for manipulating or enhancing brain physiology and behavior with neurofeedback technology. In sum, our findings suggest that context-dependent brain function and flexible behavior may only emerge from context specific CLE interactions. This supports the idea that CLE interactions may have a broad impact on our cognition (Clark, 2008) and may also shape social interactions (Froese et al., 2014). Materials and Methods Simple Conceptual Model: The time traces in Figure 1 were calculated by Euler's method, integrating the equations presented in the text with 𝑤 = −10, 𝜏 = 1, and time step 𝑑𝑡 = 0.01. Zebrafish experiment: We analyzed the calcium signal (∆F/F) at various sample frequencies (ca. 23 Hz) across 1908 cells in 32 fish, see Ahrens et al. (2012) for full description of experimental method. In addition to calcium sources (putative neurons), swim vigor was measured by taking electrical recordings from motor neurons in the fish's tail. In the original study, the gain (i.e., the multiplicative factor between fictive swim power, and the speed of visual feedback) was alternated between a high and low gain condition every 30 s. 22 We studied cells analyzed in Ahrens et al. (2012), which showed modulations of the mean ∆F/F, depending on the gain conditions. This gain alternating protocol is not relevant to the current study. To reduce this variability in data, we subtracted the mean activity level in each gain setting in our analysis (from both brain and behavior variables). Notably, our main results were qualitatively the same, even without such subtraction of the means. We analyzed data taken from a 6-min recording of 16 prominent calcium sources per fish, putative neurons, across 600 trials. In the first 3 min, the fish performed the closed-loop optomotor behavior. For the subsequent 3 min, each fish was presented with an open-loop stimulus (no longer actively controlling their environment), which is a repeat of what the animal experienced in the previous 3 min. In this period, sensory input to the animals was identical and the only difference was the absence of closed-loop dynamics, allowing us to study how reafferent signal affected neural activity. In Figure 2B, low frequency power was calculated as the mean log power of a neuron in the range of 0.010.15 Hz, averaged over all recorded neurons. Correlation was calculated as the average pairwise correlation (6589 pairs of cells were analyzed) between neurons in the same fish, averaged over all trials. Figure 2C plots the change in this pairwise correlation against the change in the log low frequency power (averaged for the pair) between the closed-loop and the replay condition. Note: for fish data, non-parametric tests were used and we did not assume normality. In addition, both results reported in Figure 2B were highly significant under a paired t-test, with p values <10-5. We fitted the data with linear filters that describe the interaction between individual neurons and the environmental variable. Note: here we define the environmental variable (E) as the activity of motor neurons, whose history uniquely determines the visual stimulus. We calculated the linear filter that minimizes the mean square error between a driven variable 𝑦(𝑡) and the convolution of a driving variable 𝑥(𝑡) and filter 𝐹(𝑡) over time. Filters were 23 constructed as a superposition of Laguere functions. We use Laguere functions up to the order that stopped the optimization process by choosing the best Akaike Information Criterion (Akaike, 1987). Almost all filters had an order that was mid-range between 1 and 15, indicating the robustness of the chosen order. We inferred two filters for each cell. We determined how the environment drives the brain, 𝐹(𝑡), by directly calculating the filter between the closed-loop environment and brain in the replay condition 𝐸 → 𝐵′, see Figure 3A (magenta line). We determined how the brain drives the environment, 𝐺(𝑡), by first calculating the residual variability of brain in the open-loop condition that cannot be accounted for by the closed-loop environment, i.e., 𝑅𝐵′(𝑡) = 𝐵′(𝑡) − 𝐹 ∗ 𝐸(𝑡) and subsequently calculating how 𝑅𝐵′(𝑡) drives the environment in the open-loop condition, effectively determining 𝐵′ → 𝐸′. Self-feedback is then straightforwardly estimated as the convolution of both filters 𝐻(𝑡) = 𝐹 ∗ 𝐺(𝑡). In Figure 2B, we calculated the low frequency power as the mean log power of a neuron in the range (0.010.15 Hz) and took the ratio between the closed-loop and replay condition. To analytically calculate this low frequency power ratio from the estimated filters, we can first write the dynamics of the brain in the closed- and open-loop conditions in the frequency domain as 𝐵(𝜔) = 𝐻(𝜔)𝐵(𝜔) + 𝜖(𝜔) = (1 − 𝐻(𝜔)) −1 𝜖(𝜔) 𝐵′(𝜔) = 𝐻(𝜔) 𝐵(𝜔) + 𝜖′(𝜔), where 𝜖(𝜔) and 𝜖′(𝜔) are the Fourier transforms of the noise in the closed- and open-loop conditions that we assume as having the same frequency spectrum. The ratio of the power between each condition is simply 𝐵(𝜔)2 𝐵′(𝜔)2 = 1 𝐻(𝜔)2+1−𝐻(𝜔)2, 24 where 𝐻(𝜔) is the estimated combined filter in the frequency domain. From this curve we can then straightforwardly calculate the mean log power of a neuron in the range (0.010.15 Hz). In Figure 3D, we calculated the 𝐸 → 𝐵′ filter in the replay condition and the 𝐸 → 𝐵 filter in the closed-loop condition as above, but using the Hermite rather than the Laguere functions to capture the acausal (t<0) side of the filter. Notably, the 𝐸 → 𝐵 filter in the closed-loop condition generally has an acausal component, because the brain 𝐵 and the environment 𝐸 are mutually interacting (see below). On the other hand, the 𝐸 → 𝐵′ filter in the replay condition is identical to the 𝐹(𝑡) filter defined in the previous paragraph (although with minor differences due to the use of the Hermite- rather than Laguere-based functions). Hence, the two filters (𝐸 → 𝐵′ and 𝐸 → 𝐵) are generally different and this difference originates from the presence of CLE feedback (𝐵 → 𝐸 interaction) in the closed-loop condition. Here, we quantify this difference using a simple linear model. In the closed-loop condition, the system is described by 𝐵(𝜔) = 𝐹(𝜔)𝐸(𝜔) + 𝑅𝐵(𝜔) 𝐸(𝜔) = 𝐺(𝜔)𝐵(𝜔) + 𝑅𝐸(𝜔) in the Fourier domain. Note that we again assumed the same interactions 𝐹(𝜔) and 𝐺(𝜔) in the closed-loop and replay conditions. Based on the first equation, the second equation is also described as 𝐸(𝜔) = (1 − 𝐻(𝜔)) −1 (𝐺(𝜔)𝑅𝐵(𝜔) + 𝑅𝐸(𝜔)). Therefore, the 𝐸 → 𝐵 filter in the closed-loop condition is 𝐵(𝜔)𝐸∗(𝜔) 𝐸(𝜔)𝐸∗(𝜔) = 𝐹(𝜔) + ( ∗ ∗ (𝜔) 𝐸(𝜔)𝑅𝐵 ) 𝐸(𝜔)𝐸∗(𝜔) = 𝐹(𝜔) + ( 𝐺(𝜔) ∗ 𝑅𝐵(𝜔)2 ) 𝐸(𝜔)2 1 − 𝐻(𝜔) where * describes complex conjugate. Hence, this filter is different from 𝐹(𝜔) by the second term. To predict the second term in the absence of knowing B, we assume 𝑅𝐵(𝜔)2 ≈ 𝑅𝐵′(𝜔)2, where the latter spectrum is based on the residual 𝑅𝐵′ computed in the replay condition. 25 Whisker model: We model a whisker comprised of two sections that are connected by hinges at the center and the base, which are constrained by muscles (simple torsion springs), see Supplementary Figure 3A. Whisking is implemented by driving the equilibrium position of the base spring. The center spring has an equilibrium value of zero angular displacement and thus tends to align both sections. A horizontal solid wall is placed above the whisker and, as the whisker collides with the wall, it deforms accordingly (Supplementary Figure 3B). By adjusting the relative stiffness of each torsion spring, we can control the degree to which the base angle is affected by contact events, e.g., if the whisker is very flexible, the base angle will change continuously, despite contact of the tip, see Supplementary Figure 3B. Note: the feedback in this model could be equally mediated by velocity or even acceleration, rather than whisker position, but neither choice would make a significant difference to our conclusions, as long as the CLE feedback constitutes a net negative feedback. We fixed the base spring, k1 = 1, and identified the stiffness of the whisker with k2. For Figures 4AE, we considered a rigid whisker, but relaxed this assumption in Figure 4F and Figures 3B and C. We simulated the dynamics of the whisker by minimizing a Lagrangian description of the configurational energy of the massless whisker under the constraint of the solid wall. We model a cortical circuit comprising N excitatory and N inhibitory neurons that interact with a single whisker. Dynamic activity of neuron i (𝑖 = 1, . . . , 𝑁 are excitatory and 𝑖 = 𝑁 + 1, . . . ,2𝑁 are inhibitory, with 𝑁 = 100) is modeled as a linear dynamical system by 𝑥𝑖 = −𝑥𝑖 + ∑ 2𝑁 𝑗=1 𝜔𝑖𝑗𝑥𝑗 − 𝑎𝑖 − 𝜔𝜃𝑦 𝜃 + 𝜉𝑖 + 𝐼 Differential equations are solved by the forward Euler integration method with time-bin dt = 0.5 ms. Hereafter, all time derivatives are taken to represent single-step differences divided by dt (e.g. 𝑥(𝑡) = [𝑥(𝑡 + 𝑑𝑡) − 𝑥(𝑡)]/𝑑𝑡), but we omit the ms time unit. 𝜔𝑖𝑗 is the synaptic strength from neuron j to i, 𝑎𝑖 is an adaptation current that produces low frequency (ca. 1 Hz) up/down-like oscillations (Compte et al., 2003; Gentet et al., 2010; Curto et al., 2009) in the 26 absence of neuron/whisker interactions, 𝜃 is whisker angle (positive: protracted and negative: retracted) interacting with neurons with weight 𝜔𝜃𝑦 = 0.002, 𝐼 is exafferent input that takes a non-zero value upon whisker stimulation, and 𝜉𝑖 is independent white noise of unit variance added to each neuron. Here, we interpret 𝑥𝑖 as both the firing rate and membrane potential, assuming a roughly linear relationship between the two. Entries in the connectivity matrix are assigned as 𝜔𝑖𝑗 = 𝑏𝑖𝑗𝐽 + 𝑏′𝑖𝑗𝑔 for excitatory synapses (𝑗 = 1, . . . , 𝑁) and 𝜔𝑖𝑗 = −𝑏′′𝑖𝑗𝑔 for inhibitory synapses (𝑗 = 𝑁 + 1, . . . ,2𝑁), where 𝑏𝑖𝑗, 𝑏′𝑖𝑗, 𝑏′′𝑖𝑗 are all random binary values that take 𝑏0 with probability 𝑝 = 0.1 and 0 with probability 1 − 𝑝, respectively. The weights are scaled by 𝐽 = 1 𝑝𝑁 and 𝑔 = 𝑔0 √2𝑁𝑝(1−𝑝) , so that dynamics are insensitive to the parameter values of 𝑝 and 𝑁. Note: all excitatory and inhibitory neurons behave similarly but adding sparse recurrent connections between randomly selected pairs neurons can account for inter-neural variability (Renart et. al 2010). The parameter 𝑔0 = 0.05 controls inter-neural variability, and a value less than 1 reproduces highly synchronized up/down-like fluctuations during the quiet state. Finally, the scaling of the connectivity matrix 𝑏0is determined such that the lead eigenvalue of the connectivity matrix is close to unity (≈ 0.975) and the dynamics are close to instability. The adaptation current is integrated as 𝑎 𝑖 = −0.07 𝑎𝑖 + 0.008𝑥𝑖 Over time, the adaptation variable slowly builds upon neural activity and suppresses neurons, resulting in the ca. 1-Hz oscillation. Consequently, in the absence of whisking, implemented by setting 𝜔𝜃𝑦 = 0, this simple network reproduces the power spectrum and cross-correlogram of neurons in the cortex (Figure 4). The whisker protraction of the base (𝜃1, see Supplementary Figure 3A) is driven by the sum of activity in the excitatory population and an external CPG activity, 𝑢; i.e., it is modeled as 𝜃1 = −0.93 𝜃1 + 𝜔𝑦𝜃 𝑁 𝑁 ∑ 𝑥𝑖 𝑖=1 + 𝑢 27 and is driven by the sum of activity of the excitatory population. 𝜔𝑦𝜃 = 0.085 describes the relative strength of the cortex vs. the CPG in driving the whisker variable. With this parameter, the whisker is mainly driven by the CPG and is modulated by cortical activity. During whisking, the whisker rhythm is imposed by the CPG. The rhythm is generated by a simple stochastic oscillator, given by 𝑢 = −.98𝑢 + 2𝜋 𝐹𝑤ℎ𝑖𝑠𝑘𝑣 + 𝜉𝑢 𝑣 = −.98𝑣 − 2𝜋 𝐹𝑤ℎ𝑖𝑠𝑘𝑢 + 𝜉𝑣, where 𝐹𝑤ℎ𝑖𝑠𝑘 = 10 Hz is the frequency of the oscillator and 𝜉𝑢 , 𝜉𝑣 are independent Gaussian white noise. Aside: In the current model, most excitatory neurons respond to whisker retraction and drive whisker protraction. Adding a separate counterpart population that responds to whisker protraction and drives whisker retraction does not change the model's behavior. In the current model, the tuning of cortical neurons to whisker position (Diamond et. al 2008) is mainly inherited from thalamocortical input. Passive whisker stimulations are simulated by injecting exafferent input 𝐼 = 0.035 to the cortical neurons for 25 ms. The magnitude of the exafferent input is selected such that it approximately matches the evoked change over the standard deviation of the membrane potential (𝛥𝑉𝑚 / 𝜎(𝑉𝑚)) in response to magnetic whisker deflection during the whisking condition (Crochet and Petersen, 2006). During active touch events, we injected exafferent input (𝐼 = 0.035) to the cortical neurons on contact, for the duration of the contact event, but for no longer than 25 ms. The model was run for 200 s in the closed loop, open-loop, and sustained period of active touch to calculate all quantitative measures. The power in Figure 4C is averaged over all neurons and a quadratic spline fitted to the data. 28 Chernoff discrimination: To calculate signal-to-noise-ratios, we calculated the Chernoff distance (Cover, 2012) between two probability distributions, 𝑝𝐼(𝑥) and 𝑝0(𝑥), in the presence or absence of a sensory event, respectively. Specifically, 𝛹(𝑝𝐼𝑝0) ≡ −𝑚𝑖𝑛 0<𝜆<1𝑙𝑜𝑔 ∫ 𝑝𝐼 𝜆(𝑥) 𝑝0 1−𝜆(𝑥) 𝑑𝑥 For our model, the probability distribution for each condition is well described by a Gaussian distribution 𝑝(𝑥) = 2𝜋𝐶−1/2𝑒𝑥𝑝(− (𝑥 − 𝜇)𝐶−1(𝑥 − 𝜇)), 1 2 where C and 𝜇 are covariance matrix and vector of means, respectively. By substituting this into the expression for Chernoff distance and employing the Gaussian integral identity and expressing the Chernoff distance in terms of 𝐶0, 𝐶𝐼 and 𝜇0, 𝜇𝐼, we calculate the covariance and mean between three neurons, randomly selected from the network described in the first section. We calculate covariances across ensembles of 500 networks every 10 ms for a period of 1 s, starting at the onset of the sensory event. Minimization with respect to 𝜆 is computed numerically. Monkey experiment: ECoG activity was recorded from 128 subdural electrodes implanted in a macaque monkey. All experimental and surgical procedures were performed in accordance with experimental protocols (No. H24-2-203(4)) approved by the RIKEN ethics committee and the recommendations of the Weatherall report, "The use of non-human primates in research". Subject: One monkey (male Macaca mulatta, aged 12 years, wild-type) was used in the experiment after magnetic resonance imaging. Before the implantation of subdural ECoG electrodes, the monkey was familiarized with the experimental settings and trained with a fixation task. During the fixation task, the monkey sat in a primate chair with the head in a fixed position using a helmet custom-made for the 29 monkey. The monkey was housed individually in the room with a 12-h lightdark cycle (lights on at 8:00 AM), and participated in the experiment in the daytime. The monkey had also participated in other experiments involving fixation and voluntary saccadic eye movements more than half a year previous to this study. Electrode Implantation: Subdural electrodes were surgically implanted after the completion of fixation training. To anesthetize the monkey, we administered ketamine (5 mg/kg, intramuscular), atropine (0.05 mg/kg), and pentobarbital (20 mg/kg, intravenous). We adjusted the dose of pentobarbital based on monkey's response to pain and heart rate. We chronically implanted a customized 128-channel ECoG electrode array in the subdural space (Unique Medical, Tokyo, Japan; Nagasaka et al., 2011). The monkey was rewarded only in trials for which fixation was maintained from FP onset until its disappearance. ECoG signals were recorded at a sampling rate of 1 kHz using a Cerebus data acquisition system (Blackrock, UT, USA) and down-sampled to 200 Hz. Behavioral variables, such as pupil size and eye position, were recorded with a custom eye-tracking system (Nagasaka et al., 2011). Visual stimuli were constructed in Pyschtoolbox and feedback implemented by streaming ECoG data in real-time into Matlab. We used LIBSVM (Chang and Lin, 2011) under Matlab to implement the SVM. During a 10-min training session, the monkey passively viewed consecutive 500-ms blocks of either vertical or horizontal gratings in each trial (total of about 200 trials). We found that the stimulus was most reliably decoded by applying SVM to inter-electrode correlations. We collected 100-ms runs of ECoG data from 50 electrodes for each stimulus condition and constructed the SVM vector by calculating the correlation between electrodes and flattening half (upper right triangle) of the resulting matrix to form a vector. We trained an SVM to classify vertical and horizontal grating on all collected vectors. 30 Classification performance strongly depended on the time interval used for the classifier, typically reaching > 90% cross-validated classification accuracy with 400-ms-long intervals. In order to reduce temporal delay for neurofeedback, we classified stimuli using short 100-ms intervals. This typically gave > 60% cross-validated classification accuracy which is still markedly higher than chance level. A better classification performance was also achieved from grating orientations separated by 90, rather than 180 degrees, motivating our choice for horizontal and vertical gratings for the neurofeedback. We then estimated the decision thresholds, by calculating a distribution of the decision values for each stimulus condition from the training data. These were roughly Gaussian with a positive and negative mean for the horizontal and vertical grating stimuli, respectively. We estimated the threshold for the decision value above (𝜃ℎ𝑜𝑟) and below (𝜃𝑣𝑒𝑟𝑡) which vertical and horizontal gratings are shown, respectively (see main text), by calculating the absolute value difference between the distributions and selecting the peaks of maximal difference. No gratings were presented for at least 1 s after the fixation point onset, to avoid presenting stimulus before the monkey fixated his eyes. During testing, we consecutively sampled 100-ms runs of ECoG data, constructed an SVM vector, and used the SVM kernel to calculate the decision variable. The Matlab processing time for classifications was shorter than 100 ms. We were able to record 50108 trials for each of the closed-loop and replay conditions per experiment. We present data from four recording days in the text. We recorded a minimum of 326 trials each under the closed-loop and replay conditions. Together with the calibration of the decoder, this added to a reasonable experimental length for the subject to maintain his attention. In addition, the number of trials was comparable to that of typical experiments performed in visual psychophysics. The power spectrum of the decision value (Figure 5B) was calculated from the ECoG signal within 10002800 ms after the fixation point onset. Trials in which the monkey failed to fixate 31 his eyes for at least 2800 ms were eliminated from the data set. The improvement in the performance was measured by the difference between the positions in each condition (Replay - Closed) relative to the average position (Replay + Closed) / 2 within each time-bin. Normality for applying the t-test (Figure 5C) was examined using the KolmogorovSmirnov goodness-of-fit hypothesis test (P > 0.07). The performance improvement for each trial pair was summarized by root-mean-square of this quantity measured over an interval of 16002100 ms. We computed the trial-to-trial correlation between the fixation performance improvement and the power spectrum of the decision value dynamics in the interval of 10002100 ms (Figures 5F and 5G): the power spectrum within this time period reflects the brain dynamics that precede the behavioral change. We also confirmed that the overall trend of the power spectrum of the decision value within this period (Supplementary Figure 5D) had not changed from that derived from the whole stimulation period shown in Figure 5B. Acknowledgements: We would like to thank the authors of Ahrens et. al. (2012) for kindly providing the zebrafish data and Poulet et. al. (2012) providing the rodent whisking data. We would like to thank Charles Yokoyama and Alexandra V Terashima for comments on the manuscript, and Misha Ahrens, Hideaki Shimazaki, Hokto Kazama, Masanori Murayama, Sylvain Crochet and L.F. Abbott for helpful discussions. Author Contributions: C.L.B and T.T. developed the theory. C.L.B. simulated the whisker model and analyzed the zebrafish data, in discussion with T.T. C.L.B. and T.T. planned the monkey experiment. C.L.B., S.T., T.Y., K.T. and Y. N. conducted the monkey experiment with the help of N.F. S.T. analyzed the monkey data based on the initial account by C.L.B. C.L.B. and T.T. wrote the manuscript. References Ahissar, E., & Kleinfeld, D. (2003). Closed-loop neuronal computations: focus on vibrissa somatosensation in rat. Cerebral Cortex, 13(1), 53-62. 32 Ahrens, M. B., Li, J. M., Orger, M. B., Robson, D. N., Schier, A. F., Engert, F., et al., (2012). Brain-wide neuronal dynamics during motor adaptation in zebrafish. Nature, 485(7399), Akaike, H. (1987). Factor analysis and AIC. Psychometrika, 52(3), 317-332. Berger, H. (1929). Über das elektrenkephalogramm des menschen. European Archives of Psychiatry and Clinical Neuroscience, 87(1), 527-570. Bruno, R. M., & Simons, D. J. (2002). Feedforward mechanisms of excitatory and inhibitory cortical receptive fields. The Journal of neuroscience, 22(24), 10966-10975. Buzsáki, G., Anastassiou, C. A., and Koch, C. (2012). The origin of extracellular fields and currents-EEG, ECoG, LFP and spikes. Nature reviews neuroscience, 13(6), 407-420. Castro-Alamancos, M. A. (2004). Absence of rapid sensory adaptation in neocortex during information processing states. Neuron, 41(3), 455-464. Chang, C., & Lin, C. (2011). LIBSVM: A library for support vector machines. ACM Transactions on Intelligent Systems and Technology (TIST), 2(3), 27. Churchland, M. M., Byron, M. Y., Cunningham, J. P., Sugrue, L. P., Cohen, M. R., Corrado, G. S., et al., (2010). Stimulus onset quenches neural variability: a widespread cortical phenomenon. Nature neuroscience, 13(3), 369-378. Clark, A. (2008). Supersizing the mind: Embodiment, action, and cognitive extension: Embodiment, action, and cognitive extension. Oxford University Press. Compte, A., Sanchez-Vives, M. V., McCormick, D. A., & Wang, X. (2003). Cellular and network mechanisms of slow oscillatory activity (< 1 Hz) and wave propagations in a cortical network model. Journal of neurophysiology, 89(5), 2707-2725. Constantinople, C. M., & Bruno, R. M. (2011). Effects and mechanisms of wakefulness on local cortical networks. Neuron, 69(6), 1061-1068. Crochet, S., & Petersen, C. C. (2006). Correlating whisker behavior with membrane potential in barrel cortex of awake mice. Nature neuroscience, 9(5), 608-610. Crochet, S., Poulet, J. F., Kremer, Y., & Petersen, C. C. (2011). Synaptic mechanisms underlying sparse coding of active touch. Neuron, 69(6), 1160-1175. Crapse, T. B., & Sommer, M. A. (2008). Corollary discharge across the animal kingdom. Nature Reviews Neuroscience, 9(8), 587-600. Curtis, J. C., & Kleinfeld, D. (2009). Phase-to-rate transformations encode touch in cortical neurons of a scanning sensorimotor system. Nature neuroscience, 12(4), 492-501. Curto, C., Sakata, S., Marguet, S., Itskov, V., & Harris, K. D. (2009). A simple model of cortical dynamics explains variability and state dependence of sensory responses in urethane-anesthetized auditory cortex. The Journal of Neuroscience, 29(34), 10600-10612. deBettencourt, T. M., Cohen, J. D., Lee, R. F., Norman, K. A., & Turk-Browne, N. B. (2015). Closed-loop training of attention with real-time brain imaging. Nature neuroscience, 18(3), 470-475. 33 Diamond, M. E., von Heimendahl, M., Knutsen, P. M., Kleinfeld, D., & Ahissar, E. (2008). 'Where'and'what'in the whisker sensorimotor system. Nature Reviews Neuroscience, 9(8), 601-612. Eliades, S. J., & Wang, X. (2008). Neural substrates of vocalization feedback monitoring in primate auditory cortex. Nature, 453(7198), 1102-1106. Eliades, S. J., & Wang, X. (2008). Neural substrates of vocalization feedback monitoring in primate auditory cortex. Nature, 453(7198), 1102-1106. Fanselow, E. E., & Nicolelis, M. A. (1999). Behavioral modulation of tactile responses in the rat somatosensory system. The Journal of neuroscience, 19(17), 7603-7616. Fee, M. S., Mitra, P. P., & Kleinfeld, D. (1997). Central versus peripheral determinants of patterned spike activity in rat vibrissa cortex during whisking. Journal of neurophysiology, 78(2), 1144-1149. Felsen, G., & Dan, Y. (2005). A natural approach to studying vision. Nature neuroscience, 8(12), 1643-1646. Froese, T., Iizuka, H., & Ikegami, T. (2014). Embodied social interaction constitutes social cognition in pairs of humans: a minimalist virtual reality experiment. Scientific reports, 4. Fry, S. N., Rohrseitz, N., Straw, A. D., & Dickinson, M. H. (2009). Visual control of flight speed in Drosophila melanogaster. Journal of Experimental Biology, 212(8), 1120-1130. Fu, Y., Tucciarone, J. M., Espinosa, J. S., Sheng, N., Darcy, D. P., Nicoll, R. A., et al., (2014). A cortical circuit for gain control by behavioral state. Cell, 156(6), 1139-1152. Gentet, L. J., Avermann, M., Matyas, F., Staiger, J. F., & Petersen, C. C. (2010). Membrane potential dynamics of GABAergic neurons in the barrel cortex of behaving mice. Neuron, 65(3), 422-435. Goard, M., & Dan, Y. (2009). Basal forebrain activation enhances cortical coding of natural scenes. Nature neuroscience, 12(11), 1444-1449. Hafed, Z. M., Lovejoy, L. P., & Krauzlis, R. J. (2011). Modulation of microsaccades in monkey during a covert visual attention task. The Journal of Neuroscience, 31(43), 15219- 15230. Harris, Kenneth D, and Alexander Thiele. "Cortical state and attention." Nature Reviews Neuroscience 12.9 (2011): 509-523. Hill, D. N., Curtis, J. C., Moore, J. D., & Kleinfeld, D. (2011). Primary motor cortex reports efferent control of vibrissa motion on multiple timescales. Neuron, 72(2), 344-356. Ito, M. (1972). Neural design of the cerebellar motor control system. Brain research, 40(1), 81-84. Aström, K. J., & Murray, R. M. (2010). Feedback systems: an introduction for scientists and engineers. Princeton university press. Kawato, M. (1999). Internal models for motor control and trajectory planning. Current opinion in neurobiology, 9(6), 718-727. 34 Keller, G. B., & Hahnloser, R. H. (2009). Neural processing of auditory feedback during vocal practice in a songbird. Nature, 457(7226), 187-190. Keller, G. B., Bonhoeffer, T., & Hübener, M. (2012). Sensorimotor mismatch signals in primary visual cortex of the behaving mouse. Neuron, 74(5), 809-815. Kinomura, S., Larsson, J., Gulyas, B., & Roland, P. E. (1996). Activation by attention of the human reticular formation and thalamic intralaminar nuclei. Science, 271(5248), 512-515. Kleinfeld, D., Ahissar, E., & Diamond, M. E. (2006). Active sensation: insights from the rodent vibrissa sensorimotor system. Current opinion in neurobiology, 16(4), 435-444. Lee, S., Carvell, G. E., & Simons, D. J. (2008). Motor modulation of afferent somatosensory circuits. Nature neuroscience, 11(12), 1430-1438. Lisberger, S. G., Morris, E., & Tychsen, L. (1987). Visual motion processing and sensory- motor integration for smooth pursuit eye movements. Annual review of neuroscience, 10(1), 97-129. Lovick, T. (1972). The behavioural repertoire of precollicular decerebrate rats. The Journal of physiology, 226(2), 4P-6P. Matyas, F., Sreenivasan, V., Marbach, F., Wacongne, C., Barsy, B., Mateo, C., et al., (2010). Motor control by sensory cortex. Science, 330(6008), 1240-1243. McGinley, M. J., David, S. V., & McCormick, D. A. (2015). Cortical membrane potential signature of optimal states for sensory signal detection. Neuron, 87(1), 179-192. Nguyen, Q., & Kleinfeld, D. (2005). Positive feedback in a brainstem tactile sensorimotor loop. Neuron, 45(3), 447-457. Niell, C. M., & Stryker, M. P. (2010). Modulation of visual responses by behavioral state in mouse visual cortex. Neuron, 65(4), 472-479. O'Regan, J. K., & Noë, A. (2001). A sensorimotor account of vision and visual consciousness. Behavioral and brain sciences, 24(05), 939-973. Otazu, G. H., Tai, L., Yang, Y., & Zador, A. M. (2009). Engaging in an auditory task suppresses responses in auditory cortex. Nature neuroscience, 12(5), 646-654. Pierret, T., Lavallée, P., & Deschênes, M. (2000). Parallel streams for the relay of vibrissal information through thalamic barreloids. The Journal of Neuroscience, 20(19), 7455-7462. Pinto, L., Goard, M. J., Estandian, D., Xu, M., Kwan, A. C., Lee, S., et al., (2013). Fast modulation of visual perception by basal forebrain cholinergic neurons. Nature neuroscience, 16(12), 1857-1863. Polack, P., Friedman, J., & Golshani, P. (2013). Cellular mechanisms of brain state- dependent gain modulation in visual cortex. Nature neuroscience, 16(9), 1331-1339. Poulet, J. F., & Petersen, C. C. (2008). Internal brain state regulates membrane potential synchrony in barrel cortex of behaving mice. Nature, 454(7206), 881-885. Poulet, J. F., Fernandez, L. M., Crochet, S., & Petersen, C. C. (2012). Thalamic control of 35 cortical states. Nature neuroscience, 15(3), 370-372. Renart, A., de la Rocha, J., Bartho, P., Hollender, L., Parga, N., Reyes, A., et al., (2010). The asynchronous state in cortical circuits. science, 327(5965), 587-590. Schneider, D. M., Nelson, A., & Mooney, R. (2014). A synaptic and circuit basis for corollary discharge in the auditory cortex. Nature. Scott, S. H. (2004). Optimal feedback control and the neural basis of volitional motor control. Nature Reviews Neuroscience, 5(7), 532-546. Semba, K., & Komisaruk, B. (1984). Neural substrates of two different rhythmical vibrissal movements in the rat. Neuroscience, 12(3), 761-774. Shibata, K., Watanabe, T., Sasaki, Y., & Kawato, M. (2011). Perceptual learning incepted by decoded fMRI neurofeedback without stimulus presentation. science, 334(6061), 1413-1415. Smith, K. U. (1962). Delayed sensory feedback and behavior. W. B. Saunders, Philadelphia Steriade, M. (2001). Impact of network activities on neuronal properties in corticothalamic systems. Journal of neurophysiology, 86(1), 1-39. Steriade, M. (2001). Impact of network activities on neuronal properties in corticothalamic systems. Journal of neurophysiology, 86(1), 1-39. Steriade, M. (2005). Cellular substrates of brain rhythms. Electroencephalography: Basic principles, clinical applications, and related fields, 5, 31-83. Steriade, M. M., & McCarley, R. W. (2013, March 9). Brainstem control of wakefulness and sleep. Springer Science & Business Media. Szwed, M., Bagdasarian, K., & Ahissar, E. (2003). Encoding of vibrissal active touch. Neuron, 40(3), 621-630.Eliades, S. J., & Wang, X. (2008). Neural substrates of vocalization feedback monitoring in primate auditory cortex. Nature, 453(7198), 1102-1106. Tan, A. Y., Chen, Y., Scholl, B., Seidemann, E., & Priebe, N. J. (2014). Sensory stimulation shifts visual cortex from synchronous to asynchronous states. Nature, 509(7499), 226-229. Urbain, N., Salin, P. A., Libourel, P., Comte, J., Gentet, L. J., & Petersen, C. C. (2015). Whisking-Related Changes in Neuronal Firing and Membrane Potential Dynamics in the Somatosensory Thalamus of Awake Mice. Cell reports, 13(4), 647-656. Von Holst, E. (1954). Relations between the central nervous system and the peripheral organs. The British Journal of Animal Behaviour, 2(3), 89-94. Vyazovskiy, V. V., Olcese, U., Hanlon, E. C., Nir, Y., Cirelli, C., & Tononi, G. (2011). Local sleep in awake rats. Nature, 472(7344), 443-447. Welker, W. (1964). Analysis of Sniffing of the Albino Rat 1). Behaviour, 22(3), 223-244. Wolf, R., Voss, A., Hein, S., Heisenberg, M., & Sullivan, G. (1992). Can a Fly Ride a Bicycle?[and Discussion]. Philosophical Transactions of the Royal Society B: Biological Sciences, 337(1281), 261-269. 36 Wolpert, D. M., & Ghahramani, Z. (2000). Computational principles of movement neuroscience. nature neuroscience, 3, 1212-1217. Y. Nagasaka, K. Shimoda, N. Fujii. Multidimensional recording (MDR) and data sharing: an ecological open research and educational platform for neuroscience. PloS one, 6(7), e22561 (2011). Zagha, E., Casale, A. E., Sachdev, R. N., McGinley, M. J., & McCormick, D. A. (2013). Motor cortex feedback influences sensory processing by modulating network state. Neuron, 79(3), 567-578. Figures Figure 1 A simple model of the brain-environment interaction. (A) A schematic description of brain-body-environment interactions during closed-loop behavior. The brain receives two types of sensory input: exafferent input that originates from the environment and reafferent input, which, while mediated by the environment, results from the consequences of an animal's own actions. (B-D) Schematic diagrams (Top) and the model's representative brain activity traces (Bottom) under the following three conditions. In the open-loop condition (B), the brain receives no reafferent input and exhibits collective activity that spontaneously fluctuates. In the closed-loop condition (C), reafferent input constitutes a CLE feedback to the brain. If this feedback is negative, the gain of the brain is reduced and fluctuations are suppressed. In the replay condition (D), the brain receives a replay of the reafferent input in 37 the closed-loop condition as exafferent input. Any differences from the closed-loop condition are caused by the absence of CLE feedback because the sensory input is identical to that in the closed-loop condition. In this condition, the gain of the brain is not suppressed and fluctuations are much larger than in the closed-loop condition with negative feedback. Figure 2. CLE feedback suppresses neural fluctuations and correlations. (A) Photograph of a paralyzed larval zebrafish (left) in the experimental setup (right), supported by pipettes that record motor activity. (B) Population averages of logarithmic low frequency power (mean over interval of [0.01 0.15] Hz) (left) and pairwise intra-neural correlations (right) were both suppressed under the closed-loop condition relative to the 38 replay condition. (C) These changes in pairwise correlations and low frequency power (replay – closed) were highly correlated in the recorded neurons. Figure 3. CLE feedback predicts suppression of neural fluctuations and correlations. (A) Dynamic interactions were estimated for each neuron by fitting linear filters, whose population averages, after normalizing to peak amplitudes, are summarized. Schematic interactions of the brain (a single neuron) and the environment (motor neuron activity) are shown under the closed-loop and replay conditions (inset). On average, the brain positively drove the environment (B'→E', purple line); the environment negatively drove the brain 39 (E→B', pink line), and, by combining these two effects, we found that self-feedback (E→B'→E', cyan line) was negative. (B) These filters were then used to predict changes of neural fluctuations under the two conditions. The predicted changes in each neuron based on the filters exhibited strong correlation with the actual changes. Some neurons (top 10%, red dots) exhibited strong negative CLE feedback and were stabilized under the closed-loop condition as predicted by our theory. (C) The location of these neurons are overlaid with the morphology of a reference zebrafish brain (colors as in B). Top panel, side view; bottom panel, top view. Neurons that have strong negative CLE feedback and are strongly stabilized were predominantly located in the cerebellum. (D) The dynamic relation of neuronal activity and motor activity for each neuron in the closed-loop condition (quantified by the E→B filter, naively computed) was qualitatively different from that in the replay condition (the E→B' filter) (inset). For the closed-loop-stabilized cells in (C, red dots), this difference could be explained well by the B'→E' filter from the replay condition (Inset: E→B, black line; E→B', blue line; E→B estimated taking into account of the B'→E' filter from the replay condition, cyan line). The degree to which the difference was accounted for by the B'→E' filter was computed for each cell as a fraction of the mean square error explained. This quantity was positively correlated with the log-power ratio, indicating that the difference in dynamic relation of neuronal activity and motor activity between the closed-loop and replay conditions was greater in the closed-loop stabilized cells. 40 Figure 4 A model of the rodent brain state transition. (A) A schematic of the model: 100 excitatory (Exc) and 100 inhibitory (Inh.) neurons receive CLE feedback via a single whisker driven by a central pattern generator (CPG). Triangle and circles represent excitatory and inhibitory synapse respectively. Onset of whisking occurs when the CPG is switched on. CLE feedback is negative overall because the neurons that elicit whisker protraction are assumed 41 to drive whisker retraction. (B) Membrane potential of cortical neurons (gray lines for individual neurons and black line for population average) and whisker position (blue line) during quiet attentive (Q) whisking (W), and periods of active touch (T). Large and synchronous fluctuations of membrane potential were suppressed during whisking. Active touch elicited reliable responses in these neurons. The vertical dotted line marks the onset of whisking and the vertical red lines mark onset of individual touch event. The power spectrum (C; inset for variance) and cross-correlation (D; inset for correlation matrix of randomly sampled neurons-color warmth indicates the degree of correlation) of membrane potential are averaged over cortical neurons and shown for each condition. Low frequency fluctuations and inter-neural correlation are suppressed during whisking but are recovered during the period of active touch. (E) Membrane potential traces for the Q, W, and T conditions. Sensory events begin at time 0. (F) The discriminability of each type of sensory event. Discrimination performance was similar under Q and W because both signal (exafferent evoked response) and noise (spontaneous fluctuations) were large under Q and both are small during W. Discrimination performance for active touch events were improved relative to the Q and W conditions unless the whisker was too flexible. Discrimination performance was improved with increasing whisker stiffness, reflecting the degree to which CLE feedback was stopped during touch events. 42 Figure 5 Real-time visual neurofeedback altered brain activity and improved fixation performance. (A) The experimental setup. An estimate of the current visual stimulus (horizontal or vertical grating) is decoded (decision value) every 100 ms from electrodes distributed across the visual cortex. Whenever decision values indicate a high confidence for either grating stimuli the opposite grating stimulus is presented. The stimulus is a gray screen for small decision values (low confidence). The box at the bottom shows the representative decision value dynamics (red trace) and sampling points (circles) used during neurofeedback. The shaded areas show the periods during which the visual stimulus were 43 presented; the dotted lines, show the decision thresholds (purple: horizontal; green: vertical). Time 0 indicates appearance of the fixation point. (B) Difference of power spectrum of the decoder decision value (Replay-Closed). Average and SEM of whole trials. (C) Summary of the 3-4 Hz amplitude of the difference of decision value in (B). (D) The improvement of fixation performance was quantified by the relative difference between the eyes deviation from fixation point (Replay-Closed) / (Replay+Closed)×2. The mean and SEM of whole trials were used for the analysis. Fixation performance was quantified by the root mean square of the distance from fixation point. (E) Average fixation performance during the time interval [1600-2100 ms] in (D). (F) Trial-to-trial correlation coefficient between the differential fixation performance and the differential amplitude for individual frequencies (bin width: 0.25 Hz). (G) Summary of correlation coefficient for 3-4Hz. The errorbar shows a bootstrapped SEM. Asterisk indicates statistical significance (*: P < 0.05; **: P < 0.005). Figure 6 A summary of the three experimental systems studied (Top). In all systems, CLE feedback plays a critical role in determining brain dynamics and behavior, as summarized in the table (Bottom). 44 SUPPLEMENTARY INFORMATION Contents S1 Supplementary fish data S2 A role for CLE feedback in the brain state transition S3 Whisker model S4 A simple conceptual model of CLE feedback S5 Decoding of visual stimulus based on ECoG signals S6 Alternate schemes for CLE feedback in the whisker system S1 Supplementary fish data 45 Figure S1 (A) The decorrelation effect is not an artifact of measurement noise. Changes in pairwise correlations and change in low frequency power were highly correlated in the recorded neurons even when the each calcium traces (both cells and motor neurons) were thresholded (the threshold was equal to the mean plus one standard deviation of the calcium signal measured over both replay and closed loop conditions) (Spearman's rank correlation =0.57, p<10-8). (B) The increase of motor activity in the closed-loop condition does not explain reduction in neural fluctuations and correlation. (B,top): Changes in mean motor activity(replay-closed) and change in low frequency power (mean over interval [0.01 0.15] Hz) per trial (low frequency power is averaged over all cells within a given trial) are positively 46 correlated (r=0.18, p<10-2, Spearman's rank correlation). B(bottom): Changes in mean motor activity and changes in pairwise correlations (averaged over all pairwise interactions in a given trial) are not significantly correlated (r=0.03, p>0.5, Spearman's rank correlation). (C) Filter describing behavioral feedback. A linear filter that describes behavioral feedback (E→E') is also strongly negative. Following the kernel method outlined in the methods section we calculated the behavioral feedback as a direct filter between the closed loop environment and environment in the replay condition (E→E'). This also indicates that the CLE feedback to the is strongly negative. Note: the magnitude of the CLE feedback is much greater than the cell self-feedback reflecting the fact that cellular variability was much greater than variability across animals. S2 A role for CLE feedback in the brain state transition In this section, we analyze if sensory input through infraorbital nerve (ION) plays a role in coordinating whisking behavior, thalamic spiking activity, and cortical local field potential (LFP). Previous results based on simultaneous recording from whisker, thalamus, and cortex exhibited that thalamic spiking rate increased and low frequency power of cortical LFP decreased during whisking behavior (Poulet et. al 2012). Here, we reanalyze this data set and quantify temporal coordination between (1) 5-20Hz power of the whisker position, denoted by Whisker; (2) thalamic spiking rate computed with 20ms averaging window, denoted by Thalamus; (3) and 1-20Hz cortical LFP power, denoted by Cortex, recorded from ION-intact animals (n=22) and ION-cut animals (n=19). Raw recordings and these processed traces are shown in Supplementary Figure 2A for an example animal. We chose the 1-20 Hz range for the analysis of cortical LFP power because notable brain-state-dependent changes were previously observed in this range (\cite{poulet14}). The spectrogram was computed using 2s window to reliably estimate the predominant 1Hz power in cortical LFP and the window was gradually shifted in 20ms steps. Next, we computed cross-correlation functions between these 3 quantities: Whisker-Thalamus, Whisker-Cortex, and Thalamus-Cortex. While the resulting cross-correlation functions were noisy in each animal, a mean cross-correlation function averaged over each animal group exhibited clear 47 common properties. In both ION-intact and ION-cut animals, whisking behavior lead correlated increase in the thalamic activity and decrease in the cortical slow oscillations. Consistent with this result, the thalamic activity was negatively correlated with the low-frequency cortical LFP fluctuations (Supplementary Figure 2B). However, the position of the mean cross-correlation peaks was significantly shifted in ION-cut animals relative to the ION-intact animals (Supplementary Figure 2B). Specifically, the peak of the Whisker-Thalamus cross-correlation was delayed for 400 ms (p=0.02, bootstrap test) and the peak of Whisker-Cortex cross-correlation was delayed for 200 ms (p=0.03, bootstrap test) in ION-cut animals. On the other hand, the temporal relationship between the thalamic spiking activity and the low- frequency cortical LFP power was not significantly altered as assessed by the Thalamus-Cortex correlation function (p>0.05, bootstrap test). The corresponding bootstrap statistics were computed by randomly resampling animals from the two groups, assuming a null hypothesis that the two animal groups are the same (see, the inset panels for the bootstrap statistics about the difference of the cross- correlation peaks). These analyses suggest that, after a whisking onset, the brain state transition was delayed in ION-cut animals relative to ION-intact animals. Thus, while sensory input is not necessary for the brain state transition, it was necessary for inducing short-latency brain state transitions. 48 Figure S2 (A) A simultaneous recording of whisker position (Top), thalamic spikes (Middle), and cortical LFP (Bottom) in an example animal (Poulet et. al 2012). .Based on these raw traces (black), brain-state-relevant quantities (red) are computed and shown in each panel: 5-20 Hz power of the whisker position (Top), thalamic spiking rate (Middle), 1-20 Hz power of the cortical LFP (Bottom). (B) A cross-correlation function between Whisker and Thalamus (Top), Whisker and Cortex (Middle), and Thalamus and Cortex (Bottom) for ION-intact animals (blue) and ION-cut animals (red), where the inset panels show the bootstrap statistics about the difference of the cross-correlation peaks. The Whisker-Thalamus and the Whisker-Cortex correlation functions were significantly shifted by the ION cut. S3 Whisker model 49 Figure S3: (A) A whisker comprising of two sections with a joint angle and base angle . The base and joint are constrained by two springs with spring constants k1, and k2 respectively. Whisking is implemented by driving the equilibrium position of the base spring. The center spring is in equilibrium at zero angular displacement and tends to align the 50 whisker sections. The whisker is length L and massless but constrained by a solid wall placed at y, where y<L. (B) The relationship between the joint (red) and base (blue) angle in a flexible, k2 =.1, (middle) and stiff, k2 =10, (bottom) whisker. (C) The dependence of cortical fluctuations and responses on whisker stiffness. Model data from quiet attentive (Q, green), whisking (W,blue) and during contact events (red) for whiskers of increasing stiffness. (C,top): average cross correlation of membrane between neurons during Q, W and period of active touch for different stiffnesses (k2/(k1=1)). (C,bottom): same as top but for low frequency power (average over the range [0.5, 2]Hz). S4 A simple conceptual model of CLE feedback We illustrate our whisker theory with an extension of simple conceptual model presented in the text, Figure 1. It describes the dynamic interaction between brain activity (here membrane potential of neurons) and the body (here whisker position) (Supplementary Figure 4). Specifically a simplified brain-environment system was modeled as a stochastic 1- dimensional first order linear ordinary differential equation as follows, 𝐵𝑐 (𝑡) = (𝜔𝑠 − 1)𝐵𝑐(𝑡)– 𝜔𝐸(𝑡) + 𝜉(𝑡) + 𝐼(𝑡) 𝐸(𝑡) = 𝐵𝑐(𝑡) integrated with a Euler step 𝑑𝑡 = 0.01, where 𝐵𝑐 is a dimensionless dynamical variable representing the collective activity of neurons (e.g. local field potential) and 𝐸 is the environmental variable (e.g. whisker position). Note: for simplicity we assume that the dynamics of the environment variable are much faster compared with the brain dynamics and, thus, E rapidly converges to current B. The negative CLE feedback robustly reduces neural fluctuations unless feedback delay is too large. The parameter, 𝜔𝑠 = 0.95, is the self- coupling within the brain, 𝜔, is the magnitude of the environmental influence on the brain, 𝐼 is external input and, 𝜉 is unit variance white noise of mean 0. An open-loop condition, i.e., the absence of environmental coupling, is modeled by setting, 𝜔 = 0 and the closed-loop condition by setting 𝜔 = 0.4. Here we have set 𝜔𝑠 to be close to criticality (𝜔𝑠 = 1) to show that even weak CLE feedback can significantly reduce neural fluctuations. However, this 51 choice is not central to our theory because strong CLE feedback works likewise when the system is away from criticality. Passive sensory stimulation was modeled by setting, 𝐼 = 1.5 for a time period of 20. Interruption of the reafferent signal was implemented by fixing the environmental variable for the same period, i.e, setting 𝐸(𝑡′) = 𝐸(𝑡) for 𝑡 < 𝑡′ < 𝑡 + 20. The replay condition is constructed by first recording the environmental variable, 𝐸, in the closed-loop condition and subsequently replaying this recording to the brain at a different time. The brain activity in the replay condition, 𝐵𝑟, is described by 𝐵𝑟 (𝑡) = (𝜔𝑠 − 1)𝐵𝑟(𝑡) − 𝜔𝐸(𝑡) + 𝜁(𝑡) where 𝜔 = 0.4 as in the closed-loop condition and 𝜁 is white noise during the replay condition. Like our simple conceptual model under these conditions CLE feedback reduces neuronal gain and suppresses neural fluctuations in the closed-loop condition (Supplementary Figure. 4E, blue lines). Furthermore, the responses of neurons to exafferent input are also suppressed (shaded black interval in Supplementary Figure. 4E). However, large responses during the active state are recovered if exafferent input coincides with brief interruptions of CLE feedback (e.g., in Supplementary Figure. 4E the environmental variable is briefly fixed -- on the surface of an object for whisker touch events -- after the onset of stimulation for a period indicated by the shaded red interval; see Supplementary Figure. 4 for a schematic diagram). Specifically, this brief interruption temporarily increases neuronal gain and enhances the brain's response to exafferent input. 52 Figure S4 CLE feedback explains the changes in brain state at the onset of active behavior. A-D: Schematics of the interaction of a brain (B) and environment (E) variable. In the open- loop condition (A), the brain receives passive exafferent input only. In the closed-loop condition (B), an environment variable mediates negative CLE feedback to the brain. In the reafferent-cut condition, (C), this feedback is briefly interrupted. In the replay condition (D), the environmental input to the brain in the closed-loop condition is recorded (denoted by Er) and played back to an identical brain (albeit with different noise) at a later time. In all conditions, in addition to the reafferent input, the brain receives a short pulse of exafferent input. (E): Traces of the brain and environment variables under different conditions. In the open loop condition (green) the brain shows large spontaneous fluctuations and large response to exafferent stimulation, which are suppressed during the closed-loop condition (red). The gray bar denotes periods of stimulation. In contrast, the brain variable exhibits large and reliable response to the same input when combined with a brief interruption of the 53 environment (the pink bar denotes the period of stimulation and reafference interruption). In the replay condition (magenta), the brain behaves similarly to the open-loop condition despite receiving the same input from the environment as the closed-loop condition. S5 Decoding of visual stimulus based on ECoG signals 54 Figure S5 Decoding of visual stimulus based on ECoG signal. (A) ECoG electrodes used in the visual neurofeedback experiment. Small black circles indicate the loci of electrodes. The electrodes surrounded by red circles were used for decoding by SVM. (B) Decision value distribution for a trained SVM, in the visual neurofeedback experiment. The purple and the green curves indicate cumulative probability distribution of the decision value during presentation of vertical and horizontal gratings, respectively. The dashed lines show the criteria for presenting feedback stimulus. (C) Averaged temporal evolution of stimulus presentation frequency within a trial during the visual neurofeedback experiment. The visual stimulus started to appear typically 1300 ms after the onset of fixation point. To avoid presenting stimulus before monkey fixated its eyes, no stimulus was presented 0-1000 ms after the appearance of fixation point. (D) Difference of power spectrum of the decoder decision value (Replay-Closed) during early period (before the end of behavioral improvement, 1000-2100 ms). The trend was not changed from the result for whole stimulus- presentation period (1000-2800 ms), including the significant reduction of the 3-4 Hz power by the neurofeedback. Convention follows Figure 4B. The black bar indicates the 3-4 Hz, used in the analysis of behavioral correlation in Figure 4f and 4G. S6 Alternate schemes for sensory feedback in the whisker system A 55 B 56 ParameniscalCortexVPMPOmFNPrVSpViRetractionThalamusBrainstemExcitation InhibitionCPGEIProtractionEfferenceReafferenceExafferenceLeminiscalParameniscal C Figure S6: Three models for whisker circuits mediating a negative CLE feedback. (A) Net activation of the modeled cortical population drives neurons in the facial nucleus (FN) to drive 57 ParameniscalCortexVPMPOmFNPrVSpViThalamusBrainstemExcitation InhibitionCPGEIProtractionEfferenceReafferenceExafferenceLeminiscalParameniscalExcitation InhibitionCortexVPMPOmFNSpViRetractionThalamusBrainstemCPGEIProtractionEfferenceReafferenceExafferencePrVTop Down ControlLeminiscalParameniscal whisker protraction. This in turn reduces excitatory sensory input to the modeled cortical population because this is driven by retraction. Here negative CLE feedback is mediated implicitly at the periphery. (B) Protraction information could be conveyed along the full pathway but net inhibitory input to the modeled cortical population result because POm inhibits the cortex. In either model circuit, the initial activation of cortical neurons causes subsequent suppression of their activity by feedback through the whisker circuit, constituting a negative sensory feedback loop. These two hypotheses are testable but not necessarily mutually exclusive. (C) An alternative hypothesis is the whisker feedback is completed in the brain stem and changes in cortical activity are driven by activity changes in the thalamus. Our whisker model remains abstract in terms of known vibrissa system anatomy and, in particular, the relay stations between the cortex and a whisker. The exact concordance of the model with known vibrissa system anatomy is beyond the scope of this paper, but we provide a more detailed to demonstrate a possible anatomical explanation of our model and provide a means for the research community to experimentally examine CLE feedback in specific biological circuits. In our model (Supplementary Figure 6A), we assume that projections between regions are largely excitatory (c.f. Ahissar 2010). Importantly, we distinguish two subcortical pathways that signal afferent input to cortical neurons - one for transmitting reafferent input and the other for transmitting exafferent input. This distinction could reflect the separation between a parameniscal pathway i.e., via thalamic POm, conveying reafferent signals, and a lemniscal pathway, i.e., via thalamic VPM, conveying exafferent input (Pierret et al., 2000; Urbain et al., 2015). Accordingly, we modeled exafferent input to cortical neurons by using a stereotypical pulse upon each whisker contact and brief deflection, and reafferent input proportional to whisker angle reflecting motor efference (Szwed et al., 2003). Regardless of how the properties of reaffererent input - whisking phase, absolute position, or their temporal 58 derivatives - are encoded by the pathway they do not change the main conclusion of our model as long as the CLE feedback constitutes net negative feedback. In agreement with the anatomy, we assume that cortically generated motor signals modulate whisking behavior by acting on the facial nucleus (FN) (Ahissar, 2010). Because whisking behavior persists after sensory denervation (Welker, 1964), cortical ablation (Semba and Komisaruk, 1984), or decerebration (Lovick, 1972), we explicitly modeled a central pattern generator (CPG) that autonomously generates whisking patterns locating exogenous to the cortical-whisker loop (Hill et al., 2011). Thus, the FN receives input from both the cortical population and CPG and moves the whisker in the reafference model. We cannot rule out other biological pathways for negative CLE feedback. For example, negative feedback can also be mediated by dominant cortical inhibition to the modeled population of cortical neurons (Supplementary Figure 6B). In agreement, thalamocortical connections strongly innervate fast spiking neurons and consequently implement strong feedforward inhibition to the cortex (Bruno and Simons 2002). Other potential models arise from heterogeneity in cortical populations. For example, negative CLE feedback can be mediated by neurons in the barrel cortex that directly drive whisker retraction with extremely short latencies (Matyas et al., 2010). Alternatively, the dominant negative feedback loop could be subcortcial, Supplementary Figure 6C. Here the dynamics of the cortex only indirectly reflects the stabilization of the thalamus. This scheme is also consistent with reduced thalamic activity during whisking (Poulet et al., 2012). It is important to note that this implementation also fundamentally relies on stabilisation of neuronal activity by negative CLE feedback. 59 At the level of the whole vibrissa system, there are likely multiple parallel and nested feedback loops, both positive and negative (Ahissar et al., 2003). However, we assume that the overall or net feedback mediated by the cortical-whisker circuit during corresponding behavior is negative in sign which we empirically demonstrate the presence of negative CLE feedback in zebrafish active sensing. These models are experimentally testable. For example, a group of neurons that encode aspects of reafferent input, such as whisker protraction, could be genetically labeled and used for anatomical tracing studies. We can moreover study the physiological role of these neurons by optogenetically silencing them during active whisking and study how brain state as well as the animal's behavior may be altered. This specific neuronal population could also be optogenetically activated and the ensuing behaviors and changes in vibrissae information processing pathways studied. 60
1304.5674
3
1304
2013-07-26T21:28:03
Generalized cable theory for neurons in complex and heterogeneous media
[ "q-bio.NC" ]
Cable theory has been developed over the last decades, usually assuming that the extracellular space around membranes is a perfect resistor. However, extracellular media may display more complex electrical properties due to various phenomena, such as polarization, ionic diffusion or capacitive effects, but their impact on cable properties is not known. In this paper, we generalize cable theory for membranes embedded in arbitrarily complex extracellular media. We outline the generalized cable equations, then consider specific cases. The simplest case is a resistive medium, in which case the equations recover the traditional cable equations. We show that for more complex media, for example in the presence of ionic diffusion, the impact on cable properties such as voltage attenuation can be significant. We illustrate this numerically always by comparing the generalized cable to the traditional cable. We conclude that the nature of intracellular and extracellular media may have a strong influence on cable filtering as well as on the passive integrative properties of neurons.
q-bio.NC
q-bio
Generalized cable theory for neurons in complex and heterogeneous media Claude Bedard and Alain Destexhe Unit´e de Neuroscience, Information et Complexit´e (UNIC), CNRS, Gif-sur-Yvette, France (Dated: October 30, 2018) Cable theory has been developed over the last decades, usually assuming that the extracellular space around membranes is a perfect resistor. However, extracellular media may display more complex electrical properties due to various phenomena, such as polarization, ionic diffusion or capacitive effects, but their impact on cable properties is not known. In this paper, we generalize cable theory for membranes embedded in arbitrarily complex extracellular media. We outline the generalized cable equations, then consider specific cases. The simplest case is a resistive medium, in which case the equations recover the traditional cable equations. We show that for more complex media, for example in the presence of ionic diffusion, the impact on cable properties such as voltage attenuation can be significant. We illustrate this numerically always by comparing the generalized cable to the traditional cable. We conclude that the nature of intracellular and extracellular media may have a strong influence on cable filtering as well as on the passive integrative properties of neurons. PACS numbers: I. INTRODUCTION Cable theory, initially developed by Rall [1], is one of the most significant contributions of theoretical neuroscience and has been extremely useful to explain a large range of phenomena (reviewed in [2]). However, cable theory makes a number of assumptions, one of which is that the extracellular space around neurons can be modeled by a resistance, or in other words, that the medium around neurons is resistive or ohmic. While some measurements seem to confirm this assumption [3], other measurements revealed a marked frequency dependence of the extracellular resistivity [4, 5], which indicates that the medium is non-resistive. Indirect measurements of the extracellular impedance also show evidence for deviations from resistivity [6 -- 9], which could be explained by the influence of ionic diffusion [10]. Despite such evidence for non-resistive media, the possible impact on cable properties has not been evaluated. The effect of non-resistive media can be investigated by integrating this effect in the impedance of the extracellular medium, Ze, and in particular, through its frequency dependence. For example, it can be shown that Ze ∼ 1/ω for capacitive effects or electric polarization [11], Ze ∼ 1/ √ω for ionic diffusion (also called the "Warburg impedance" [12]), while Ze would be constant for a perfectly resistive medium. To integrate such effects in a given formalism, such as the genesis of extracellular potentials, our approach has been to integrate a general frequency-dependent function Ze(ω) in the formalism, and then consider specific cases [10, 12]. In the present paper, we follow this approach and generalize cable equations for media with arbitrarily complex frequency-dependent impedance. With numerical simulations, we consider specific cases such as resistive media, ionic diffusion, capacitive media, etc. We evaluate a number of possible consequences on the variation of the membrane poten- tial along the cable, and how such effects could be measured experimentally. II. METHODS All simulations were done using MATLAB. To simulate the cable structure of the models, a classic compartmental model strategy was used for simulations (see Fig. 3F), but was different from the one used in common simulator programs such as NEURON (Hines and Carnevale, [13]). Each cylindric compartment is connected to intracellular and extracellular resistances or impedances, and these are normally used to solve the cable equations. In the present paper, we used another, where Vm and ii are respectively equivalent method which consists of defining an auxiliary impedance, given by Za = Vm ii Bedard&Destexhe(October30,2018) 2 the transmembrane potential and the axial current per unit length at the point where Za is connected (see Fig. 1). This auxiliary impedance allows to take into account the influence of other compartments, including the soma, over the axial current and transmembrane potential. It is mathematically equivalent to consider the continuity conditions on axial current and transmembrane potential. The electric and geometric parameters are considered constant in each compartment, but are allowed to vary between compartments. In these conditions, Vm and ii are solution of partial differential equations (cable equations) and thus depend on spatial coordinates. FIG. 1: (Color online) Convention used to calculate the input impedance and transfer function. A cable segment of length l is repre- sented, with an impedance Za in series, at the end of the cable. This "auxiliary impedance" Za takes into account the influence of the other compartments on the axial current ii and transmembrane potential Vm in a compact form. Za = Vm(l) ii(l) where ii(l) is the current per unit length and Vm(l) is the transmembrane voltage at coordinate x = l. The cable equations simulated in this article are generalized to allow one to include media with complex electrical properties. We have designed a MATLAB code that simulates such generalized cable structures, using different types of linear density of complex impedances ([Ω/m]) and specific impedances ([Ω.m]) in each compartment. See Results for details of this method. All computations were made in Fourier space. We have applied the theory to four different types of media to evidence their effect on the spatial and frequency profile of the membrane potential. These models are called SC, FC, FO and NIC, respectively (see Table I). The SC model is the "standard model" as defined by Tuckwell [14]; the FC model corresponds to a model similar to the standard model (based on a closed circuit), but the cytoplasm and extracellular media impedances can be frequency-dependent. The FO type model is the same, with an open circuit (no return current). The NIC model includes a non-ideal capacitance similar to a previous study [15]. See Results for details of these models. All numerical simulations were made using a "continuous ball-and-stick" model, consisting of a single cylindric com- partment, described as a continuum (see Results), and a spherical soma. The dendritic compartment has a radius of 2 µm and a membrane time constant of 5 ms, which corresponds to typical values of in vivo conditions. The has a radius of 7.5 nm and the specific capacitance was of 0.01 F/m2. These parameters represent typical values used in a number of previous studies [2, 14, 16, 17]. III. RESULTS We start by generalizing the cable equations for membranes embedded within extracellular media of arbitrarily complex electrical properties. Next, we consider a few specific cases and numerical simulations. A. Generalized cable equations In this section, we redefine the cable equations taking into account the presence of complex and/or heterogeneous prop- erties of extracellular and intracellular media. Because electrically complex or heterogeneous media can display charge accumulation, one cannot apply the usual (free-charge) current conservation law. One needs to use a more general con- Bedard&Destexhe(October30,2018) 3 T ypes Model z(m) e SC Standard cable (closed-circuit) z(m) e = − rmre (ri+re)(1+iωτm) FC Frequency-dependent cable (closed-circuit) z(m) e = − rmze (zi+ze)(1+iωτm) λ2 rm ri+re rm zi+ze λ = 1+iωτm κ2 λ2 (ri+re)(1+iωτm) rm (zi+ze)(1+iωτm) rm FO Frequency-dependent cable z(m) e (open-circuit) rm zi [1 + z(m) e rm (1 + iωτm)] zi(1+iωτm) z(m) e rm (1+iωτm)] rm[1+ NIC Non-ideal cable (closed-circuit) z(m) e = − ω2rm τmτM [1+iω(τm+τM)][1+iωτm] rm zi [ (1+iωτm)(1+iωτM) 1+iω(τm+τM) ] zi rm [1 + i ωτm 1+iωτM ] TABLE I: Summary of dendritic cable types and parameters. The table gives the parameters z(m) , λ2 and κλ for different model types. e The standard model (SC) is the cable model as given by Rall, Koch and Tuckwell [1, 14, 17]. The "frequency-dependent model" (FC) correspond to a standard cable (closed circuit), but where the parameters zi and ze are allowed to be frequency dependent. In the "frequency-dependent open-circuit model" (FO), the current in the extracellular medium is "perpendicular" to the membrane (see Fig. 6). The "non-ideal cable" model (NIC) is similar to the standard model, but the capacitance of the membrane is non-ideal, as developed previously [15]. zi (see Eq. 10) and ze are respectively the impedance per unit length of the cytoplasm and of the extracellular medium, respectively, for FC type models. We write ri and re when the parameters zs do not depend on frequency (SC type model). The parameter z(m) e (see Eq. 18) is used in FO type models. servation law based on the generalized current. In Section III A 1 below, we derive this generalized current conservation law, while in Section III A 2, we use this generalized conservation law to derive the generalized cable equations. 1. Generalized current conservation law in heterogeneous media In this section, central to our theory, we show that the free-charge current conservation law (~j f ) does not apply to systems with complex electrical properties. Another, more general, conservation law must be used, the generalized current conservation law. We derive here the conservation law for the membrane current in arbitrarily complex media, starting from first principles. Maxwell theory of electromagnetism postulates that the following relation is always valid for any medium: ∇ × ~H = ~j f + ∂ ~D ∂t , (1) where ~H is the magnetic field, and ~j f is the current density of free charges, and ∂ ~D ∂t is the displacement current density. We define the generalized current density ~j g as: ~j g = ~j f + ∂ ~D ∂t = ~j f + ~j d , (2) where ~j d is the displacement current density. It is important to note that the term ∂ ~D field varies in time). ∂t = εo ∂ ~E ∂t is different from zero, even in the vacuum (assuming that the electric Bedard&Destexhe(October30,2018) 4 The interest of using the generalized current, is that it is always conserved in any given volume, for any type of medium, as we explain below (see also Appendix IV). In the case of an electric field in a homogeneous and locally neutral medium, we have ∇ · ~j f = − ∂ρ = 0 because there cannot be charge accumulation anywhere. Because the relation ∇ · ~j g = 0 applies to any type of medium, we also have ∇ · ( ∂ ~D ∂t ) = 0. Thus, in a homogeneous locally-neutral medium, we have two independent current conservation laws: one law applies to the free-charge current i f and another one applies to the displacement current i d . Note that in a homogeneous medium i d is not necessarily negligible, but the application of the current conservation law on i f can be done independently of the existence of i d because the two laws are independent. f ∂t This is the framework assumed in the standard cable theory, in which the extracellular medium is resistive and homo- geneous, the displacement current i d is negligible, and there cannot be charge accumulation inside the dendrites nor in the extracellular medium. We will see below that these assumptions do not hold for complex extracellular media. If the medium is heterogeneous, then charge accumulation will necessarily appear in the presence of an applied electric field. Capacitive effects is an example of such charge accumulation. In such a case, the two current conservation laws on i f and i d do not apply to every region of space (see Appendix B). However, the generalized current conservation on i g is still valid in all cases. Thus, to derive cable equations in heterogeneous media, one must use the generalized current conservation law, as done in the next section. 2. Application of the generalized conservation law to cable equations To start, we consider a small portion of membrane surface and build a domain in the intracellular side, which is limited by the interior surface of the membrane, while the other surfaces of the domain are located inside the cytoplasm (see Fig. 2). FIG. 2: (Color online) Definition of a domain D inside the cytoplasm and adjacent to the membrane. Due to conductance variations in the membrane, or due to charged currents, the total charge in domain D varies. One cannot consider that the displacement current across the surface of domain D is zero, because this would be in contradiction with Maxwell-Gauss law (see Appendix B). Black circles represent negative charges on the interior surface of the membrane, as well as in the cytoplasm, while blue circles indicate positive charges at the exterior side of the membrane. Bedard&Destexhe(October30,2018) 5 Using such a definition, in resting conditions, the intracellular side has an excess of negative charges, which are adjacent to the membrane. In such a state, we can calculate the free charge density in this domain from Maxwell-Gauss law: ~D · n dS = cst , Q(t) = (cid:9) ∂D where ∂D is the surface of the considered domain[34] Now, suppose that a conductance variation occurs in the domain (for example following the opening of an ion channel). This will induce a charged current in domain D and therefore, there will be a variation of the total charge included within domain D, which implies a non-zero displacement current i d across the surface ∂D surrounding domain D (without this current, the system would be in contradiction with Maxwell-Gauss law; see Appendix B). (3) In such conditions, we have: i d = dQ dt = (cid:9) ∂D ∂ ~D ∂t · n dS , 0 (4) Can we neglect this current to study the variations of the membrane potential along the cable? Because it is difficult to give a rigorous answer to this question [18, 19], in particular when i d is non-zero, we consider the generalized current i g because this current is conserved independently of i d (see previous section). This will allow us to treat cable equations without making any hypothesis about charge accumulation inside or outside of the cable. Moreover, to stay as general as possible, we include a frequency and space dependence of the electric parameters, which will allow us to simulate the effect of media of different electric properties, such as capacitive or diffusive [10 -- 12]. In this context, the linking equations must be expressed in their most general form [10]: −∞ ~D(~x, t) = R +∞ ~j f (~x, t) = R +∞ −∞  εi(~x, t − τ) ~E(x, τ) dτ [σe i (~x, t − τ) ~E(~x, τ) According to this scheme, the generalized current density ~j g i inside the cytoplasm obeys: i (~x, t) = Z +∞ ~j g −∞ [σe i (~x, t − τ) ~E(~x, τ) + εi(~x, t − τ) ∂ ~E ∂t (~x, τ)] dτ (5) (6) where σe i (~x, t) is the intracellular electric conductivity function and εi(~x, t) is the intracellular electric permittivity function. The first term in the integral accounts for energy dissipation phenomena, such as calorific dissipation (Ohm's differential law) and diffusion phenomena. The second term represents the effect of charge density variations in the volume elements. In Fourier frequency space, Eq. 6 becomes algebraic. g i (~x, ω) = [σe ~ i (~x, ω) + iωεi(~x, ω)] ~E(~x, ω) (7) Moreover, we have ∇ × ~E = 0, which implies ~E = −∇V because electromagnetic induction is negligible in biological tissue (in the absence of magnetic stimulation[35]). If we now consider a one-dimensional cylindric cable of constant radius a (Fig. 3A), the generalized current at a position x of the cable can be written as: i g i (x, ω) = ~j g i (x, ω) · (πa2 n) = −πa2[σe i (x, ω) + iωεi(x, ω)] · ∂Vi ∂x (x, ω) (8) where Vi is the intracellular voltage difference with respect to a given reference (which can be far away). In the following of the text, we will call "compartment" a cylindric cable with constant radius and with uniform electric parameters (see Fig. 1). It is important to note that this compartment does not need to be isopotential, and the membrane potential will depend on the position on the compartment (see scheme in Fig. 1). Bedard&Destexhe(October30,2018) 6 FIG. 3: (Color online) Compartments and equivalent electrical circuits of the membrane and cable segments. A and C depict different configurations in a cable of constant diameter, with their respective equivalent electrical circuits shown in B and D. E is the equivalent electrical circuit of a membrane compartment of the cable, and F is the equivalent circuit obtained for three compartments. Vi is the intracellular potential relative to the reference, Ve is the extracellular potential relative to the same reference, zi is the cytoplasm impedance, rm/dx and z(m) e /dx are respectively the impedances of ion channels and the input impedance of the extracellular medium as seen by the transmembrane current. ie is the output current of a cable element in the extracellular medium, and ii is the axial current. The membrane potential Vm j equals Vi j − Ve j and may vary according to the position x j. If we assume that the impedance (per unit length) of cytoplasm zi can be expressed as: then the axial current can be written as: zi = 1 πa2[σe i (x, ω) + iωεi(x, ω)] , i g i (x, ω) = − 1 zi ∂V i ∂x (x, ω) (9) (10) This expression is similar to the traditional cable equation [2, 14, 16], with the exception that the parameter zi is complex (with units of [Ω/m])[36]. In addition, the transmembrane current i⊥m over a cable length dx can be expressed as: i⊥m(x, t) = im (x, t) dx = 2πadx [ Cm ∂Vm(x, t) ∂t + σe m e (Vm(x, t) − Em) ] = dx [ cm ∂Vm(x, t) ∂t + (Vm(x, t) − Em) rm ] (11) where Vm is the transmembrane voltage, Em is the resting membrane potential, Cm is the specific membrane capacitance (in F/m2), cm is the membrane capacitance per unit length (in F/m), σe m is the electric conductivity (in S /m), 1/rm is the Bedard&Destexhe(October30,2018) 7 linear density of membrane conductance (in S /m), e is the membrane thickness (in m) and im is the transmembrane current per unit length (in A/m) (Figs 3C-E). Applying the inverse Fourier transform, we obtain: i⊥m(x, 0) = im (x, 0) dx = 1 rm [Vm(x, 0) − 2πEmδ(0)] dx ω = 0 i⊥m(x, ω) = im (x, ω) dx = [iωcm + 1 rm ]Vm(x, ω) dx ω , 0 (12)  Note that we assume here that the resting membrane potential Em does not depend on time nor on position in the cable. Thus, we can see that the Fourier transform of Eq 10 generates a Dirac delta function for null frequency. In the following of the text, we consider frequencies different from zero, because the zero-frequency component of i⊥m is zero for a signal of finite duration, which is always the case in reality. In the model above, the expression of the transmembrane current is identical to the generalized membrane current for frequencies different from zero. In this case, the generalized current is given by: i g m = A · j g m = −2πa dx(σe m + iωεm)∇V = 2πa dx(σe m + iωεm) Vm l = dx ( 1 rm + iωcm)Vm = dx im = i⊥m (13) where l is the membrane thickness and A is the membrane surface. Assuming that the charge variations inside the channels is negligible, then the generalized current conservation law can apply to point B in the equivalent scheme (see Fig. 3 C), and we can write i g i (x + dx, ω) = i g i (x, ω) − i⊥m(x, ω) i (x, ω) − i g m (x, ω) = i g It follows that: Using Eqs. 6 and 11, we obtain: di g i (x, ω) = ∂i g i ∂x dx = −i⊥m(x, ω) = −im (x, ω) dx πa2 ∂ ∂x [(σe i (x, ω) + iωεi(x, ω)) ∂Vi ∂x (x, ω)] = [iωcm + 1 rm ]Vm(x, ω) Applying the partial derivative on the lefthand term, and dividing by πa2(σe i + iωεi), one obtains: ∂2Vi ∂x2 + 1 (σe i + iωεi) ∂(σe i + iωεi) ∂x ∂Vi ∂x · = 1 πa2(σe i + iωεi) [iωcm + 1 rm ]Vm = ziim (14) (15) (16) Note that if the righthand term was zero, then this equation would be identical to the equation describing the electric ∂x in one dimension, in which case the right potential outside of the sources [10, 12, 20], because the ∇ operator equals ex would be equal to ∇2Vi + ∇γi · ∇Vi where γi = σe i + iωεi. γi ∂ We can simplify Eq. 14 if the cytoplasm is quasi-homogeneous (assuming the scale considered is large compared to inhomogeneities due to subcellular organelles), in which case we can consider that the electric parameters of the cytoplasm are independent of position x: σe i (ω) and εi(x, ω) = εi(ω). This leads to the following expression: i (x, ω) = σe 1 zi ∂2V i ∂x2 (x, ω) = 1 πa2(σe i + iωεi) ∂2V i ∂x2 (x, ω) = [iωcm + 1 rm ]Vm(x, ω) (17) in Fourier space. If we now assume that the extracellular medium can also be considered as homogeneous (which will be valid at scales larger than the typical size of the cellular elements), then we can model the variations of the membrane potential caused by the transmembrane current i⊥m. We can model this effect from the notion of impedance, without making any hypothesis on the current field in the extracellular medium. In this case, one can associate to each cable segment dx the specific impedance of the extracellular medium, z(m) , as seen by the transmembrane current. z(m) has a similar physical meaning e e as rm, except that it is a complex number in general. In Section III B, we will see that z(m) e depends on the direction of the current field in the extracellular medium. Bedard&Destexhe(October30,2018) Without any loss of generality, we can write in Fourier space: Vi(x, ω) = Vm(x, ω) + z(m) e (ω)im(x, ω) . By substituting this last expression in Eq. 15, we obtain rm zi [1 + z(m) e rm (1 + iωτm)] ∂2Vm ∂x2 = [1 + iωτm]Vm where τm = rmcm. Thus, we can write the system in a form similar to the standard cable equation: where 8 (18) (19) (20) λ2 ∂2Vm(x, ω) ∂x2 = κ2Vm(x, ω) λ2 = rm zi · [1 + z(m) e rm (1 + iωτm)] = rm ¯zi κ2 = 1 + iωτm  for a cylindric compartment (see Eq. C8 in Appendix C). It follows that the general solution of this equation in Fourier space ω , 0 is given by: Vm(x, ω) = A+(ω)e κ(l−x) λ + A−(ω)e −κ(l−x) λ (21) for each cylindric compartment of length l and with constant diameter (see Fig. 1 for a definition of coordinates). For a given frequency, we have a second order differential equation with constant coefficients. In general, one can apply Eq. 21 for different cylindric compartments, as in Fig. 3F. In this case, one must adjust the ∂Vm ∂x (see Eq. C4 in different compartments to their specific limit conditions (continuity of Vm and of the current i g Appendix C). i = − 1 ¯zi Note that Eq. 21 is exact for a cylindric compartment of constant diameter. Thus, it is possible to use this property to simulate exactly the full cylindric compartment as a continuum with no need of spatial discretization into segments, as usually done in numerical simulators. This is only possible if the cylindric compartment has a constant diameter. This leads to an efficient method to simulate the cable equations. We will refer to this approach as "continuous compartment" in the following. As mentioned above, the mathematical forms of Eqs. 19 and 21 are identical to that of the standard cable model, but with different definitions of λ. Thus, we directly see that the nature of the extracellular medium will change the value of these parameters, which become frequency dependent. In particular, we see from Eq. 19 that changing these parameters will impact on the spatial profile of the variations of Vm, if the frequency dependence of the ratio κλ = κ λ is affected by the nature of the medium. Thus, experimental measurement of the spatial variations of Vm will be able to identify effects of the extracellular impedance only if the ratio κλ is affected. In the next section, we derive expressions to calculate the input impedance Zin(P) = Vm(P,ω) ii(P,ω) and the transfer function of Vm(Pb,ω) Vm(Pa,ω) between two positions in the cable (as a function of the ratio κλ). Later in the transmembrane voltage FT (ω) = Section III B, we will see that it is necessary to know these quantities to calculate the spatial variation of Vm and compare the standard model with the cable model embedded into complex extracellular media. 3. Method to solve the generalized cable In this section, we present the theoretical expressions which will allow us to calculate the input impedances needed for computing the membrane voltage on a cable with varying diameter. We consider the input impedance of the membrane, Bedard&Destexhe(October30,2018) 9 as well as the impedance of the extracellular medium, both of which are needed to calculate the spatial profile of the Vm in a given cable segment. We proceed according to the following steps: 1. In the previous section, we saw that it is necessary to calculate the ratio Zin(P) = source, to calculate the Vm produced at that point. Vm(P,ω) ii(P,ω) at the position of the current FIG. 4: (Color online) Branching cables. The panels A and B respectively represent a branched cable where a dendrite separates into two daughter branches, and its equivalent electrical circuit. The equivalent impedance of segment 1 is equal to the input impedances of segments 2 and 3 (zout 2 and zout 3 ) taken in parallel. One strategy is, in a first step, to separate the cable into a series of continuous compartments of constant diameter, where parameters a (Eq. 8), zi (Eq. 10), rm (Eq. 11) and z(m) (Eq 18) are constant and specific to each compartment. In a e Vm(0) second step, one calculates the (transmembrane) input impedance Zn+1 ii(0) at the begin of each compartment by taking into account the auxiliary impedance at the end of this compartment, Za = Zn+1 in (see Fig. 1) if there is no branching point. At the branching points, the auxiliary impedances are simply equal to the equivalent input impedance of n dendritic branches in parallel (where n is the number of "daughter" branches; see Fig. 4). Thus, because the input impedance at one end is equal to the input impedance of the other compartment connected to this end, one obtains a recursive relation (see Eq. C9 in Appendix C): Vm(ln+1) ii(ln+1) = Zn in = out = Zn+1 in [Zn in] = ¯zin κλn (κλnZn (κλnZn in + ¯zin) e2κλn ln + (κλnZn in + ¯zin) e2κλn ln − (κλnZn in − ¯zin) in − ¯zin) (22) where Thus, we can write ¯zi = 1 + z(m) e rm zi (1 + iωτm) Zn+1 in = F [Zn in ; ¯zin , κλn , ln] This leads to the following expression to relate the first to the nth segment: in = F [...F [F [Z1 Zn+1 in ; ¯zi1 , κλ1 , l1]; ¯zi2 , κλ2 , l2]...; ¯zin , κλn, ln] (23) Note that this algorithm is a generalization of that used to calculate the equivalent resistance for resistances in series. Indeed, for resistance in series we have req = F(...F(r1; r2); rn) where F(ra; rb) = ra + rb. The difference between this recurrence function and that of Eq. 25 essentially comes from the fact that there is no current leak in a resistance, while there is one in a dendritic compartment. 2. To calculate the profile of Vm along the cable, one must use the spatial transfer function Vm(Pn+1,ω) Vm(Pn,ω) on a continuous cylindric compartment of arbitrary length, and calculate the product of the transfer functions between each connected compartment. This leads to (see Appendix D and Eq. D3): FT (l, ω; Zn out) = κλZn out κλZn out cosh(κλl) + ¯zisinh(κλl) (24) Bedard&Destexhe(October30,2018) Vm(Pn, ω) Vm(P1, ω) = n−1 Yi=1 Vm(Pi+1, ω) Vm(Pi, ω) 10 (25) 3. To evaluate zproximal we must calculate the first impedance Z1 in which enters the recursive relation 24. This impedance corresponds to the impedance of the soma, which is given by: Z1 in = Zs + Zcs, (26) where Zs is the soma membrane impedance and Zcs is the cytoplasm impedance inside the soma. This relation is obtained under the hypothesis that the soma is isopotential, and the application of the generalized current conservation law implies where Vi and Ve are the electric potentials at both sides of the membrane, inside and outside, i g = respectively relative to a reference located far-away. Zs+Zcs ≈ Vm Vi−Ve Zs+Zcs The impedance of the bilipidic membrane is approximated by a parallel RC circuit where R = Rm is the resistance and τm = RmCm is the membrane time constant. Thus, Z1 in can be written as: Z1 in = Zs + Zcs = Rm 1 + iωτm + Zcs (27) Finally, to evaluate zdistal, we use the "sealed end" boundary condition Z1 coth(κλ1l1) (see Eq. 22). In the case of a single dendritic branch, we can write: in = ∞. In this condition, we have Z2 in = ¯zi1 κλ1 Zdistal in = ¯zi κλ coth(κλl) , (28) where l is the total length of the cable. In the next section, we turn to numerical simulations to investigate passive cable properties in the presence of complex media. We consider the most general case, where both the impedance of the extracellular medium and that of cytoplasm can be frequency dependent, and determine the respective impact on the spatial profile and frequency content of the transmembrane voltage at the level of the proximal and distal ends of the cable. B. Numerical simulations The goal of the numerical simulations is here to show how the physical nature of extracellular and intracellular media can influence the spatial and frequency profiles of the transmembrane potential. We present simulations of a "continuous ball and stick" model, which consists of a continuous cylindric compartment (described by Eq. 21), connected to a spherical soma. In this case, the impedance Za of the continuous cylindric compartment is the soma impedance (see Fig. 1). We do not investigate here the effect of complex dendritic structures, which is left for future studies. Note that what we call a "continuous cylindric compartment" actually represents an infinite number of compartments each represented by a resistance in series with a parallel RC circuit (see Fig. 3F). In a first step, we list the different types of models of intracellular and extracellular media that were used. In a second step, we present the results of numerical simulations. 1. Different types of cable models We now explain the parameters used for the simulations of the cable presented in Section III C. Because the cable equation (Eq. 19) is completely determined by the value of κλ for a given frequency, the spatial and frequency profiles of the transmembrane voltage are completely determined if the geometry and boundary conditions are Bedard&Destexhe(October30,2018) set. And because κλ is a function of 4 parameters (rm, τm, zi, z(m) )(Eq. 20) for a given frequency, we have a four-dimensional e parameter space where the two last parameters (zi, z(m) ) can be frequency dependent. We will limit our exploration of this e parameter space by only varying the physical nature of these impedances for realistic values of rm and τm, because the influence of these parameters has been largely characterized in previous studies [1, 16, 17]. Furthermore, with τm and ω fixed, the relation κλ = depends only on λ, and thus, like the classic studies on cable equations, we will use this parameter as a main determinant of the cable properties. 1+iωτm 11 λ We will explore the generalized cable equations by considering several typical cases: Closed−circuit model Open−circuit model iR Re iz ze ze FIG. 5: (Color online) Two different cable models for neurons. Left: Closed-circuit model. This is the standard cable model which forms a closed system (all inward and outward currents are balanced) and can be described by an equivalent circuit (bottom; shown here for a two-compartment model; Re and Ri are the extracellular and intracellular resistances, respectively). In this model, the current flows parallel to the neuron. Right: Open-circuit model. In this more general model, the current is allowed to flow between neighboring neurons, or between the neuron and extracellular space, with no necessary condition of local balance (top). In this case, the neuron is modeled by an open circuit (bottom), and the current flows "perpendicular" to the membrane. The equivalent circuit is modeled more generally with impedances (Ze extracellular, Zi intracellular) Standard cable model The first type of model that we will consider is the "standard cable model" (model SC in Table I), identical to that considered by Rall, Koch and Tuckwell [1, 14, 17]. In this model, the neuron is a closed system, where the inward and outward currents are balanced, forming a closed circuit (see Fig. 6, left). The extracellular current flows parallel to the dendrite, as noted previously [14]. This model is equivalent to consider that the field produced by the neuron corresponds to an electric dipole configuration. In addition, this model considers that the extracellular medium is resistive, or in other words, that the extracellular impedance is a constant. In this standard model, the extracellular impedance z(m) e is either zero (no extracellular resistivity) as in Rall's and Koch's Bedard&Destexhe(October30,2018) 12 formulations [1, 2, 17], or is equal to a constant, which is equivalent to model the extracellular medium by a resistance, as in other formulations [14, 21]. Besides its physical non-sense (the extracellular medium considered as a supraconductor), using a zero-resistance is usually justified from the fact that the extracellular resistivity is much smaller than the membrane impedance. We will see that this justification does not hold if the medium is frequency dependent, in which case for some frequency range the extracellular resistivity may be determinant. Thus, to obtain the general expression of λ and κλ for the standard model, we set z(m) (ri+re)(1+iωτm) in Eq. 20 (see Table I). rmre e = − 60 50 40 30 20 10 0 0 B 250 500 ν (Hz) 750 1,000 ) g e d ( ] λ κ [ Φ 250 500 ν (Hz) 750 1,000 C 150 x 104 2.5 A ) 2 − m ( λ κ 2 1.5 1 0.5 0 0 T F 1 0.8 0.6 0.4 0.2 0 100 5 5 50 100 50 5 100 50 150 150 150 10−4 L (m) 10−3 FIG. 6: (Color online) Spatial and frequency profile of the membrane potential in the cable model with resistive media. A and B respectively show the modulus κλ and the phase Φ[κλ] of κλ as a function of frequency ν for a continuous ball-and-stick model. C. Modulus of the transfer function FT as a function of distance L in the dendritic compartment, for frequencies equal to 5, 50, 100 and 150 Hz (see corresponding frequencies in A and B). The blue curves in − · − correspond to a standard cable model (FC, closed-circuit), with ri = 28 × 109 Ω/m and re = 18 × 109 Ω/m . The red curves correspond to the same model but in an open-circuit configuration (FO model), with ri = 28 × 109 Ω/m and z(m) e = 0.01 τm/2πaCm = 0.4 × 103 Ω.m. The black curves in −− show a non-ideal cable (NIC) model with τM = 0.01τm, ri = 28 × 109 Ω/m and re = 0 Ω/m. Frequency-dependent cable model The second type of model is an extension of the standard model, where the intra- cellular and extracellular impedances (zi and z(m) , respectively) are allowed to depend on frequency. This "frequency- e dependent cable model" (model FC in Table I) can account for example for a neuron embedded in capacitive or diffusive[37] extracellular media, or if the intracellular medium has such properties, or both. In such cases, the appropriate frequency-dependent profiles for the impedances must be used. In this frequency-dependent model, if τm is fixed, the quantity ze + zi completely determines the spatial and frequency profiles of the Vm, and how they deviate from the standard model (see Table I). To explore the effect of the impedances zi + ze, we consider three typical cases: "resistive", "capacitive" (which is in fact resistive and capacitive in parallel) and "diffusive" (which is equivalent to a Warburg type impedance). Such impedances have also been considered in previous studies [8, 12, 22]. Bedard&Destexhe(October30,2018) 13 Note that, in order to simulate the standard model, one must necessarily assume that the real part of z(m) e is negative[38], is not a passive impedance per unit length, but is active, and thus requires a source of energy, as which implies that z(m) e pointed previously [23, 24]. This point will be further considered in the Discussion. A B ) 2 − m ( λ κ 3000 2000 1000 0 T F 1 0.9 0.8 0.7 0 ) g e d ( ] λ κ [ Φ 200 400 600 800 1000 ν (Hz) C 50 0 −50 −100 5 200 400 600 800 1000 ν (Hz) 50 5 150 100 5 50 100 150 50 100 150 1 2 3 L (m) 4 5 6 x 10−4 FIG. 7: (Color online) Spatial and frequency profiles of the membrane potential for a model with resistive extracellular medium and diffusive cytoplasm. A and B: modulus κλ and phase Φ[κλ] of κλ as a function of frequency, for a continuous ball-and-stick model. C. Modulus of the transfer function FT as a function of distance for different frequencies (same arrangement as Fig. 6). The red curves correspond to a model with zero extracellular resistance. The blue curves (− · −) show models with open-circuit configuration (FO model with z(m) (1+i) √w Ω/m). The black curves (−−) show the same model with closed-circuit configuration with a resistive extracellular medium (FC model re = 18 × 109 Ω/m). Note that for the FC model, FT progressively increases from 5 to 50 Hz, then decreases between 50 and 100 Hz. e = 0.5 τm/2πaCm = 20 × 103 Ω.m) and diffusive cytoplasm (zi = 28×109 Open-circuit model In a third type of model, the "Open-circuit" model (FO in Table I), we use a different approach. Instead of considering the neuron as a closed system, where all outward currents must return to the neuron, we make no hypothesis about the return currents, and allow for example that neighboring neurons exchange currents[39]. In this case, one does not need to describe each neuron by a closed circuit, but all neurons are open circuits are are connected together (through the extracellular space). Figure 6 shows the current fluxes of the two models, the standard model is a closed circuit where the outward currents loop into the inward currents (Fig. 6A), while in the open-circuit model, all currents are exchanged with the surrounding medium (Fig. 6B). These two models correspond to different equivalent circuits (Fig. 11 in Appendix E). Note that the Open-circuit cable model is practically equivalent to the traditional (closed-circuit) cable model for an isolated neuron, if the impedance of the extracellular medium is negligible compared to the membrane impedance. Indeed, if z(m) e and ze tend to 0, then we have (see Table I): λ2 FO = lim z(m) e →0 rm zi λ2 FC = lim ze→0 (29) Similar to the frequency-dependent cable model, we will consider the three types of impedances discussed above (resistive, capacitive and diffusive) in the simulations of the Open-circuit model. In this case, we separately consider the two quantities zi and z(m) because these two parameters directly determine the value of λ in models of FO type (see e Table I). Note that in the Open-circuit model, the real part of z(m) is always positive, so there is no need of any additional e energy source (see Discussion). Non-ideal cable model The fourth type of model considered here is the "non-ideal cable model" introduced previ- ously [15]. This model postulated that the membrane capacitance is non-ideal, through the use of an additional resistance Bedard&Destexhe(October30,2018) 14 ) 2 − m ( λ κ 2 1.5 1 0.5 0 x 104 A ) g e d ( ] λ κ [ Φ 200 400 600 ν (Hz) 800 1000 C 60 40 20 0 −20 B 200 400 600 ν (Hz) 800 1000 T F 10−0.1 10−0.9 0 1 2 150 4 3 L (m) 5 50 100 5 6 x 10−4 FIG. 8: (Color online) Spatial and frequency profiles for a model with resistive cytoplasm and diffusive extracellular medium. Same arrangement of panels as for Figs. 6 and 7, but for different media. The black curves (−−) show the behavior of a closed-circuit (FC) type model with resistive cytoplasm (ri = 28 × 109 Ω/m) and diffusive extracellular space with Warburg impedance (ze = 18×109 (1+i) √w Ω/m). The red curves correspond to a closed-circuit (FC) type model with zi = 28 × 109 Ω/m and ze = 0 Ω/m. The blue curves (− · −) correspond to an open-circuit (FO) type model (ri = 28 × 109 Ω/m, z(m) (1+i) √w = 20×103 (1+i) √w Ω.m). e = τm 2πaCm 0.5 A B ) 2 − m ( λ κ 3000 2000 1000 20 0 T F 1 0.9 0.8 0.7 0 ) g e d ( ] λ κ [ Φ 200 400 600 800 1000 ν (Hz) C 1 2 3 L (m) 40 20 0 −20 −40 200 400 600 800 1000 ν (Hz) 100 150 50 5 50 100 150 5 5 4 6 x 10−4 FIG. 9: (Color online) Spatial and frequency profiles for fully diffusive cable models. Same arrangement of panels as for Figs. 6 -- 8, but using a continuous ball-and-stick model where both cytoplasmic and extracellular impedances are of diffusive (Warburg) type. The black curves (−−) correspond to a closed-circuit (FC) type model with zi = 28×109 (1+i) √w Ω/m. The red curves correspond to a closed-circuit (FC) type model with zi = 28×109 (1+i) √w Ω/m and ze = 0 Ω/m. The blue curves (− · −) correspond to a closed- circuit (FO) type model with zi = 28×109 e = 20×103 (1+i) √w Ω.m. Note that for both types of models (FO and FC), FT increases between 5 and 50 Hz, then decreases between 50 and 100 Hz. (1+i) √w Ω/m and ze = 18×109 (1+i) √w Ω/m and z(m) at the arms of the capacitor; this resistance models the fact that there is some inertia time to charge movement (or equiva- lently, a friction). Such a non-ideal capacitance resulted in a shallower frequency scaling, that is a higher capacity of the dendritic tree to propagate high-frequency events [15]. Note that in this model, the extracellular medium is modeled as a Bedard&Destexhe(October30,2018) 15 resistance, so in this respect, the non-ideal cable model is equivalent to the standard model. Mathematically, the non-ideal cable appears through the use of z(m) (see Table I), which can therefore be viewed as a particular case of an influence of the e extracellular medium on cable properties. Indeed, the non-ideal cable can be shown to be equivalent to -- or a particular case of -- the open-circuit model, where the Vm corresponds to Vi with a far-away reference (see Appendix E). We keep this model here for comparison. C. Simulation of the different models In this section, we present the results of numerical simulations of the models presented in the previous section (see Methods). The goal of these simulations is not to be exhaustive in considering all possible combinations of models, but present a few typical configurations. The central question is whether the nature of the extracellular medium can have determinant impact on cable properties, and for what type of configuration or parameter values does it happen ? ) 2 − m ( λ κ 2000 1500 5 6 8 10 1000 20 40 500 5 6 8 10 20 40 20 2 2 3 4 3 4 ) g e d ( ] λ κ [ Φ 40 60 ν (Hz) 80 100 120 25 20 15 10 5 0 −5 −10 −15 −20 −25 40 2 200 400 600 ν (Hz) 800 1000 FIG. 10: (Color online) Parameter κλ as a function of frequency for fully diffusive models. The black curves −− correspond to FO and the red curves to FC type models with a time constant of τm = 2, 3, 4, 5, 6, 8, 10, 20 and 40 ms . The FC type model was with zi = 28×109 (1+i) √w Ω.m (see Table II for the corresponding resonance frequencies). (1+i) √w Ω/m . For the FO type model, zi = 28×109 (1+i) √w Ω/m and ze = 18×109 (1+i) √w Ω/m and z(m) e = τm 2πaCm 0.5 (1+i) √w = 20×103 Analysis of the spatial profiles of Vm variations In this section, we investigate analytically and numerically different particular cases of extracellular and intracellular media to determine how the nature of these media affects the spatial and frequency profile of the membrane potential. We consider the transfer functions as defined in Table I. The analyses presented here are limited to a ball-and-stick model, which allows a better interpretation of the effect of the physical nature of the different media. The effect of complex dendritic tree morphology will be the subject of a future study. To compare the results from the different models, we have considered models with identical geometry (see Methods for parameters). Resistive models We first considered the "standard model" with resistive intracellular and extracellular media, as well as the non-ideal cable model [15]. In Figure 7, we can see that the nature of the cable model (closed-circuit or open- circuit; non-ideal) influences the modulus and the phase of κλ, as well as the spatial profile of the transfer function FT. The modulus of the transfer function depends more strongly on frequency in the FC model compared to the two other cases (Fig.7C), as observed previously [15]. Note that the parameters of the FO and NIC models were chosen such that they are equivalent (see Appendix B). Capacitive models Next, we considered models where the cytoplasm and extracellular medium are both of capacitive (RC-circuit) type. Note that we considered capacitive effects without ionic diffusion, because if both are combined, the Bedard&Destexhe(October30,2018) 16 resulting impedance is of Warburg type. This type of model will be considered next. With purely capacitive media, we observed effects very similar to the resistive model shown in Fig. 7, with slight differences only visible for large frequencies (greater than about 200 Hz (not shown). The small dimension of organelles (<< 1 µm2) within cells, as well as the distance between neighboring cells (∼ 30 nm on average) [25, 26] imply that the capacitance values of the media are necessary small compared to the membrane capacitance, and thus the purely capacitive effects (without diffusion) are likely to be negligible. τm (ms) νr (Hz) 2 3 4 5 6 8 10 20 83 54 40 30 25 20 18 8 TABLE II: Resonance frequencies of fully diffusive models for different membrane time constants. The resonance frequencies of κλ as a function of the membrane time constant τm (see Fig. 10). Resistive models with diffusive cytoplasm We next considered models where the extracellular medium was resistive as above, but where the intracellular medium (cytoplasm) was diffusive, and described by a Warburg impedance. Figure 7 shows the spatial and frequency behavior of this model. We can see that the open-circuit (FO) model shows less attenuation with distance compared to the closed-circuit (FC) model. Note that these two models give opposite variations when the extracellular medium has a zero resistance: in FC type models, FT attenuates more steeply as a function of distance when the extracellular impedance increases, whereas in FO type models, the attenuation becomes less steep. However, the spatial profile of FT also attenuates less with a diffusive cytoplasm compared to a resistive cytoplasm. The latter result is expected, because the higher the frequency the more the impedance "short-cuts" the membrane in this case. Note that the Warburg impedance used in all diffusive models considered here was applied for frequencies larger than 5 Hz. It is interesting to note that in the FC model, a resonance appears around 24 Hz in the modulus of the transfer function κλ (Fig. 7A). In contrast, the FO model does not display a resonance. Resistive cytoplasm with diffusive extracellular medium Next, we considered the opposite configuration as previously, namely a resistive model for the cytoplasm, but a diffusive extracellular medium. Three sets of parameters were chosen for the extracellular space. First, a FO type model with a resistive cytoplasm and a diffusive extracellular medium described by a Warburg type impedance (black curve in Fig. 9), and second, a FC type model with similar parameters (blue curve in Fig. 8). These two models can be justified if one takes into account the Debye layer at the edge of the membrane [8 -- 10]). The case with a zero extracellular resistance (short-cut) is also shown for comparison (red curve in Fig. 9). The latter model represents the same limit case for both FO and FC models, and therefore constitutes the frontier between the two families of curves. Fully diffusive cable models Next, we have considered the case where both intracellular and extracellular media are diffusive. Figure 9 shows the frequency and spatial profiles of the Vm for such fully diffusive models. Taking the FO model with low extracellular impedance (ze = re at 1 Hz) leads to large differences with the FC model (Fig. 9, black) compared to the FO model (blue) or the FC model with zero extracellular resistance (red). We can see that, in FC type models, the larger ze, the steeper the transfer function attenuates with distance. In contrast, in FO type models, larger zm e lead to less attenuation. This paradoxical result can be explained as follows: in FC models, ze plays as similar role as zi, such that for large values of their real part, thermal diffusion attenuates the signal; in FO models, zm e limit the leak membrane current, reducing the attenuation with distance. Thus, for large zm e , the dendrites become more "democratic" in the sense that the effect of a given input will be less dependent on its position on the dendrite. This is only the case for FO models, however. e plays a similar role as rm, and large values of zm As above, the model with zero extracellular resistance represents the same limit case for both FO and FC models, and therefore constitutes the frontier between the two models. Bedard&Destexhe(October30,2018) 17 ∂x2 = κλVm, so that the quantity ∂2Vm Resonances with diffusive models One interesting finding is that resonances appear in several models using diffusive extracellular impedances (Figs. 7 and 9). This type of resonance was studied further in Fig. 10, where one can see that a resonance in κλ also implies a resonance in FT: The Vm still attenuates with distance independently of the frequency, so that we always have ∂Vm∂x < 0. In addition, Eq. 19 shows that ∂2Vm ∂x2 increases when κλ increases with frequency, which implies that ∂Vm∂x becomes more negative because this derivative is always negative. It follows that FT attenuates more steeply with distance when κλ increases with frequency. Using a similar reasoning, one can show that FT attenuates less steeply with distance when κλ diminishes with frequency. We conclude that the rate of variation of FT with frequency is always opposed to that of κλ. Consequently, the resonance frequency must be the same for κλ and dFTd f = 0 because we have dFT d f ≤ 0 when dκλd f ≥ 0 and dFT d f ≥ 0 when dκλd f ≤ 0. We also see that the peak frequency of the resonance continuously depends on the membrane time constant (not shown). For example, for τm = 5 ms, the resonance is at about 24 Hz, and for τm = 20 ms, the resonance is at about 8 Hz (for more details see Fig. 10 and Table II). It is interesting to note that we have observed resonances only in FC type models with resistive extracellular media and diffusive cytoplasm (see Fig. 7), but resonances are present in the two types of models (FO and FC) when they are fully diffusive. IV. DISCUSSION In this paper, we have introduced a generalization of cable equations to membranes within media with complex or heterogeneous electrical properties. We have shown that generalized cable equations can treat a number of problems presently not treatable by the traditional cable equations. We have shown that the nature of the extracellular medium has a significant influence on fundamental neuronal properties, such as voltage attenuation with distance, and the spectral profile of the transmembrane potential. We enumerate below the consequences and predictions of this work, as well as outline directions for future studies. A first main result of this paper is to generalize cable equations to describe membranes in complex and heterogeneous media. To solve this problem, we have introduced the concept of generalized current, and show that the generalized current is conserved in all situations. This stands in contrast with the free-charge current, which is conserved only in special cases. For example, if the medium is electrically non-homogeneous (with conductive and non-conductive domains), there will be charge accumulation and non-conservation of the free-charge current. Thus the traditional cable formalism, which is based on the free-charge current, cannot treat this problem. With the generalized current, however, this problem can be treated in a physically plausible way, in accordance with Maxwell equations. One drawback of generalized cable equations is that they cannot be solved with available neural simulation environ- ments, such as NEURON [13], which implements the traditional cable formalism. Consequently, we have developed a specific method for the numerical simulation of generalized cables. This method is implementable with traditional simu- lation programs, such as MATLAB. Further work would be needed to determine if generalized cable equations could be included in neural simulators, as a special case. Note that specialized models different from the standard model were introduced relatively recently [22, 27] to include aspects which cannot be treated by the standard model. In [22], the cytoplasm was considered as non-resistive but capacitive, and was modeled by a RC circuit. It was estimated that this capacitive aspect is important to understand the nature of thermal noise in thin dendritic branches. [27] considers the case of the interaction between closely located dendritic branches. In this case, the authors study the phenomenon of surface polarization (see also [11]) and evaluate the magnitude of the Maxwell-Wagner time of the effective impedance of the extracellular medium, needed to have significant influences over the attenuation profile of the Vm. These two studies show that the physical nature of the intracellular or extracellular media can have significant influences on cable properties. However, they do represent very particular cases, which motivated the present study where we have attempted to consider a broad range of cases, including both intracellular and extracellular media, as well as ionic diffusion, which was not treated previously. Thus, the present study generalizes those prior studies. A second main result of this paper was to also generalize the electrical circuit representing neuronal membranes. Instead of considering the neuron as a closed system, where all outward currents return to the neuron, we have considered the Bedard&Destexhe(October30,2018) 18 more general case which allows current exchange between neighboring neurons, and thus each is represented by an open circuit. We have systematically compared open-circuit (FO) models with the traditional closed-circuit (FC) models, and found some important differences. FO models have a transfer function that depends much less on frequency and space, compared to FC models (see Figs. 7 and 8). We also showed that a previously introduced model of non-ideal cable [15] is equivalent to a traditional cable with appropriately scaled extracellular resistances (for frequencies smaller than 100 Hz; see Figs. 7 and 8 in [15], as well as the discussion in that paper). One of the most important result of this paper is the finding that the nature of extracellular or intracellular media can have a strong impact on cable properties such as voltage attenuation with distance. We have observed that the nature of the extracellular medium has an opposite impact on distance attenuation on FO and FC models. In FO models, larger extracellular impedances lead to less attenuation and electrotonically more compact dendrites. The attenuation can be remarkably diminished for fully resistive FO models, with only a few percent attenuation (Fig. 9), whereas for FC type models, the opposite was seen, the dendrites become more compact for low extracellular impedances. We can say that in these cases, the effect of distal inputs is close to that of proximal inputs, and thus the dendrite is more "democratic". It may be that this remarkable property is present in some types of neurons to reduce the attenuation of distal inputs, which constitutes another interesting direction to explore in future work. Another interesting observation is that diffusive extracellular impedances can give rise to resonance frequencies (see Figs. 7 and 9), which also appears as a resonance in κλ (Fig. 10). The resonant frequency depends on the membrane time constant, and is in the range of 5-40 Hz, which is well within the frequency range of brain oscillations such as theta, alpha, beta or gamma rhythms [28]. It is therefore possible that this resonance plays a role in the genesis of oscillatory activity by single neurons. Interestingly, we observed that the input impedance of the extracellular medium (z(m) e ) must necessarily be negative in the standard model where the medium is resistive. In a closed-circuit configuration, this means that one must neces- sarily assume a source of energy, such as an electromotive force. This source of energy can be simulated by a negative impedance. This important point was pointed in previous work, where it was called "anomalous impedance" [23, 24]. Interestingly, this constraint disappears in the open-circuit configuration. If the current field is open in the extracellular medium, then it is not necessary to assume that z(m) e is negative, and there is no need of such a source of energy. Finally, while our analysis shows that the nature of the extracellular or intracellular media may be influential on single- neuron behavior, we can also foresee consequences at the network level. First, the resonance found for some of the media may introduce a bias in the genesis of oscillatory behavior by populations of neurons. The fact that the resonance frequency only depends on membrane parameters, but not on structural parameters such as cell size, suggests that dif- ferent neurons in the network will have the same resonance frequency. It is thus conceivable that population oscillatory activity may occur at this resonance frequency. Second, the fact that the diffusive properties of media were found to be particularly impactful on the attenuation of distal inputs suggests that any regulation of these properties could have drastic consequences at the network level. If diffusive properties are modified, for example by glial cells who are known to regulate extracellular ionic concentrations [29, 30], it may affect the voltage attenuation of all cells in the network and therefore change network behavior. In conclusion, we think that the generalized cable equations allow one to treat the problem of how neuronal membranes behave in complex extracellular and heterogeneous media. Given the possible strong impact of such media as found here, future studies should evaluate in more depth whether such media are indeed influential. A possible approach would be to find "signatures" of the extracellular medium from the power spectral density of experimentally observable variables, such as the membrane potential (for a related approach, see [8]). The direct measurement of the extracellular impedance, at present bound to contradictory experimental results [3 -- 5], should give a definite indication whether the generalized cable is a necessary approach to accurately model neurons. Bedard&Destexhe(October30,2018) 19 Appendices Appendix A: Generalized current and charge conservation In this appendix, we derive the charge conservation laws for different definitions of currents (see Eqs. 1 and 2). Consider a domain D delimited by a closed surface ∂D. If we assume that the medium and the field are sufficiently regular, then the divergence theorem applies in D, and we have: (cid:9) ∂D ∇ × ~H · n dS ≡ $ D ∇ · (∇ × ~H) dv ≡ 0 because the following equality always applies: ∇ · (∇ × ~H) ≡ 0 [40] From Eqs. 1, 2 and A1, we have the following identity: (cid:9) ∂D ~j g · n dS ≡ $ ∇ · ~j g dv = 0 , D (A1) (A2) which is valid for an arbitrary domain D. One can distinguish three different types of current, the generalized current i g, the current due to free charges i f , and the displacement current i d. These currents can be defined across an arbitrary surface S, according to: i g def def = !S = !S = !S i f i d def  ~j g · n dS ~j f · n dS ∂ ~D ∂t · n dS (A3) Within these definitions, we can write that the generalized current i g is conserved at every time and independently of the nature of the medium. At every time, the inward current entering a given domain D is always equal to the outward current exiting that domain, independently of the homogeneous or heterogeneous nature of the medium. It is also independent of the fact that there may be charge accumulation in some elements of volume, because Eq. A2 always applies. Note that this generalized current conservation law does not express anything new on a physical point of view, but is the charge conservation law expressed as a function of currents. Indeed, taking into account Maxwell-Gauss law (∇· ~D = ρ f ), the definition of ~j g (Eq. 2) and the identity given by Eq. A2, we obtain the differential charge conservation law: ∇ · ~j g = ∇ · ~j f + ∇ · ∂ ~D ∂t = ∇ · ~j f + ∂ ∇ · ~D ∂t = ∇ · ~j f + f ∂ρ ∂t = 0 (A4) Appendix B: Displacement current, free current and charge accumulation In this appendix, we show explicitly that the displacement current i d can be used to formally calculate the charge variation in a given domain D. Moreover, we show that the displacement current across a closed surface ∂S which surrounds a given domain D is zero when there is no charge variation inside the domain. Bedard&Destexhe(October30,2018) By definition, the density of displacement current (Eq. A3) in frequency space is given by: ~j d(~x, ω) = iωε(~x, ω) ~E(~x, ω) where ω = 2π f . By applying the divergence on ~j d and taking into account Maxwell-Gauss law, we obtain: ∇ · ~j d = iω∇ · (ε ~E) = iωρ f 20 (B1) (B2) Thus, we can calculate the amount of free charges in a given domain D from the density of displacement current in frequency space. To do this, we have Q f (ω) = $ D ρ f (~x, ω) dv = 1 iω $ D ∇ · ~j d(~x, ω) dv ≡ ~j d · n dS = i d(ω) iω 1 iω (cid:9) ∂D (B3) where i d is the displacement current flowing across surface ∂S. Applying the inverse Fourier transform, we obtain the rate of free charge variation in domain D: f dQ dt (t) = i d(t) (B4) Therefore, one can say that the charge in the considered volume does not vary if the displacement current across surface ∂D is zero. Finally, because the differential conservation law for free charges implies: f dQ dt (t) = $ D ∂ρ f (~x, t) ∂t dv = −$ D ∇ · ~j f (~x, t) dv ≡ −(cid:9) ∂D ~j f · n dS = −i f (t) , we can then write: i g(t) = i d(t) + i f (t) = 0 (B5) (B6) when the surface is closed and when the free charge conservation law applies. Thus, the generalized current entering a given closed surface ∂D is always equal at every time to the generalized current exiting ∂D, even if there is free charge accumulation inside ∂D. However, this equality does not allow one to deduce if there are variations of free charge density inside ∂D, because the displacement current must necessary be zero across ∂D to have dQ = 0 (see Eq. B4). In other words, it is necessary that the displacement current entering ∂D is equal to the dt displacement current exiting ∂D to have a constant charge inside ∂D. Note that in any given circuit, Kirchhoff's current law always applies to the generalized current, even if there is charge accumulation inside the circuit, whereas it applies to the free charge current only assuming there is no charge accumulation inside the circuit. f Appendix C: Input impedance of a cable segment in series with an arbitrarily complex impedance In this appendix, we calculate the input impedance of a cable segment of length l when this segment is connected to an arbitrary impedance Za (see Fig. 4). By definition, we have in x = 0: Zl in[Za] = Vm (0, ω) i g i (0, ω) Applying Eq. 21 allows us to directly express Vm as a function of the cable parameters. We have Vm(0, ω) = A+(ω) eκλl + A−(ω) e−κλl, Similarly, applying Eqs. 10, 13 and 18, we obtain: i g i = − 1 zi [1 + z(m) e rm ((1 + iωτm)] ∂Vm ∂x (C1) (C2) Bedard&Destexhe(October30,2018) This last expression allows us to express the current at coordinate x = 0 as a function of the cable parameters: where i g i (0, ω) = κλ ¯zi [A+(ω) eκλl − A−(ω) e−κλl] ¯zi = 1 + z(m) e rm zi (1 + iωτm) Thus, the expression for the input impedance Zl in is given by: Zl in[Za] = ¯zi κλ · ( A+ A− ) · e2κλl + 1 ( A+ A− ) · e2κλl − 1 21 (C3) (C4) (C5) We can then evaluate the ratio A+ Applying Eqs. 21 and 10 to that point gives: A− by using the conditions of continuity of the current and of the voltage at point x = l. Thus, we have and we can write Vm(l, ω) = A+(ω) + A−(ω) (a) i g i (l, ω) = κλ ¯zi [A+(ω) − A−(ω)] (b)  Za = Vm(l, ω) i g i (l, ω) = κλ ¯zi A+ A− + 1 [ A+ A− − 1] , A+ A− = κλZa + ¯zi κλZa − ¯zi It follows that the input impedance Zl in is given by: where Zl in[Za] = ¯zi κλ (κλZa + ¯zi) e2κλl + (κλZa − ¯zi) (κλZa + ¯zi) e2κλl − (κλZa − ¯zi) ¯zi = 1 + z(m) e rm zi (1 + iωτm) (C6) (C7) (C8) (C9) Note that Zl in[Za] → ¯zi κλ when l → ∞, and Zl in[Za] → ¯zi κλ coth(κλl) when Za → ∞. Appendix D: Calculation of the transfer function FT In this appendix, we calculate the transfer function FT (l, ω; Za) = for Appendix C. Applying Eq. C5a gives: Vm(l,ω) Vm(0,ω) using the same conditions and conventions as Thus, we have Vm(0, ω) = A+(ω) eκλl + A−(ω) e−κλl Vm(l, ω) = A+(ω) + A−(ω)  FT (l, ω; Za) = A+(ω) + A−(ω) A+(ω) eκλl + A−(ω) e−κλl (D1) (D2) Bedard&Destexhe(October30,2018) Applying Eq. C7 gives the transfer function: where FT (l, ω; Za) = κλZa κλZa cosh(κλl) + ¯zisinh(κλl) ¯zi = 1 + z(m) e rm zi (1 + iωτm) Note that FT (l, ω; 0) = 0 and FT (l, ω; ∞) = 1 cosh(κλl) . 22 (D3) FIG. 11: (Color online) Equivalence of the electrical circuits of open-circuit and non-ideal cable models. The circuits A and B are equivalent when the ratio V12(ω) I(ω) of the voltage difference between points 1 and 2 and the input current between these points is invariant, and when the correspondence between the elements of these circuits are independent of frequency. Note that the values of the elements between the two circuits are related by a transformation law which is independent of frequency; this equivalence also applies to the temporal domain. In other words, according to this equivalence, the two circuits are equivalent when it is impossible to distinguish their topology from external measurements. The circuit A corresponds to the non-ideal capacitance model introduced previously [15], while circuit B corresponds to a "standard cable model" with a short-cut (zero extracellular resistivity). Appendix E: A new interpretation of the non-ideal cable In this appendix, we show that the non-ideal capacitance model introduced previously [15] is equivalent to an open- circuit resistive model if we assume that the circuits A and B in Fig. 11 are linked by the following transformation: ra = rm − rmrsc rm+rsc rm = ra + rb rb = rmrsc rm+rsc rsc = rb + r2 b ra ca = (rm+rsc) rm rsc rm− rm +rsc cm cm = r2 a (ra+rb)2 ca (E1)  We show that the Vm in the non-ideal cable model corresponds to Vi in an open-circuit (FO) type resistive model, with a reference located far-away. According to circuits A and B in Fig. 11, we have: circuit A (rsc ⊕ cm) k rm circuit B (ra k ca) ⊕ rb  (E2) Bedard&Destexhe(October30,2018) It follows that the impedances of circuits A and B are equal if we have: V12(ω) i(ω) = rm + iωrmrsccm 1 + iωcm(rm + rsc) = ra + rb + iωcararb 1 + iωcara 23 (E3) We see that the ratio V12(ω) is a homographic transform of variable ω. Consequently, ∀ω we have the following relation i(ω) V12(ω) i(ω) = aA+bAω 1+dAω = aB+bBω 1+dBω when the two circuits are equivalent. The only way to guarantee that the equivalence is independent of frequency is to assume that the corresponding coefficient of the transformations are equal. We can thus set aA = aB, bA = bB and dA = dB. This gives us 3 equations which link the 3 parameters of circuit A to those of circuit B. The solution is the transformation law (Eqs. E1). Thus, on a physical point of view, one cannot distinguish the topology of circuits A and B if we would perform external measurements. Moreover, because the functions rm = fm(ra, rb.rc), rsc = fsc(ra, rb.rc), rcm = fcm(ra, rb.rc) do not depend on frequency, their equivalence will be also valid for all frequencies. We can deduce that the two circuits will behave identically as a function of time . It follows that the Vm (between points 1 and 2) in circuit A (non-ideal capacitance) corresponds to the Vi relative to a far- away reference in circuit B (see Table I). Therefore, a model with non-ideal capacitance and zero extracellular resistance should produce a Vm equivalent to the Vi of a model with ideal capacitance and resistive extracellular medium. Thus, the frequency-scaling behavior of the Vm obtained in a previous non-ideal cable model [15] also applies to the resistive FO model, but only if one studies the intracellular potential Vi. Acknowledgments Research supported by the CNRS, the ANR (Complex-V1 project) and the European Union (BrainScales FP7-269921 and Magnetrodes FP7-600730). [1] Rall, W. (1962) Electrophysiology of a dendritic neuron model. Biophys J. 2: 145-167. [2] Rall, W. (1995) The theoretical foundations of dendritic function. MIT Press, Cambridge, MA. [3] Logothetis N.K., Kayser, C. and Oeltermann, A. (2007) In vivo measurement of cortical impedance spectrum in monkeys : implications for signal propagation. Neuron 55: 809-823. [4] Gabriel, S., Lau, R.W. and Gabriel, C. (1996a) The dielectric properties of biological tissues : I. Literature survey. Phys. Med. Biol.. 41 : 2231-2249. [5] Gabriel, S., Lau, R.W. and Gabriel, C. (1996b) The dielectric properties of biological tissues : II. Measurements in the frequency range 10 Hz to 20 GHz. Phys. Med. Biol.. 41: 2251-2269. [6] B´edard, C., H. Kroger, and A. Destexhe. 2006. Does the 1/f frequency scaling of brain signals reflect self-organized critical states ? Physical Review Lett. 97: 118102. [7] Bazhenov M, Lonjers P, Skorheim P, Bedard C and Destexhe A. (2011) Non-homogeneous extracellular resistivity affects the current-source density profiles of up-down state oscillations Phil Trans R Soc A 369: 3802-3819. [8] B´edard, C., Rodrigues, S., Roy, N., Contreras, D. and Destexhe, A. (2010) Evidence for frequency-dependent extracellular impedance from the transfer function between extracellular and intracellular potentials. J. Computational Neurosci. 29: 389-403. [9] Dehghani, N., B´edard, C., Cash, S.S., Halgren, E. and Destexhe, A. (2010) Comparative power spectral analysis of simultane- ous elecroencephalographic and magnetoencephalographic recordings in humans suggests non-resistive extracellular media. J. Computational Neurosci. 29: 405-421. [10] B´edard, C. and Destexhe, A. (2011) A generalized theory for current-source density analysis in brain tissue. Physical Review E 84: 041909. [11] B´edard, C., Kroger, H. and Destexhe, A., (2006b) Model of low-pass filtering of local field potentials in brain tissue. Phys. Rev. E 73:051911. [12] B´edard, C. and Destexhe, A. (2009) Macroscopic models of local field potentials and the apparent 1/f noise in brain activity. Biophys. J. 96: 2589-2603. [13] Hines, M.L. and Carnevale, N.T (1997) The NEURON simulation environment. Neural Computation 9: 1179-1209. [14] Tuckwell, H.C. (1988) Introduction to Theoretical Neurobiology: Linear cable theory and dendritic structure. Cambridge Uni- versity Press, Cambdridge, UK. Bedard&Destexhe(October30,2018) 24 [15] B´edard, C. and Destexhe, A. (2008) A modified cable formalism for modeling neuronal membranes at high frequencies. Biophys. J. 94: 1133-1143. [16] Johnston, D. and Wu, S.M. (1995) Foundations of cellular neurophysiology. MIT Press, Cambridge, MA. [17] Koch, C. (1999) Biophysics of Computation. Oxford University press, Oxford, UK. [18] Destexhe, A. and B´edard, C. (2012) Do neurons generate monopolar current sources ? J. Neurophysiol. 108: 953-955. [19] Riera, J.J., Ogawa, T., Goto, T., Sumiyoshi, A., Nonaka, H., Evans, A., Miyakawa, H. and Kawashima, R. (2012) Pitfalls in the dipolar model for the neocortical EEG sources. J. Neurophysiol 108: 956-975. [20] B´edard, C., Kroger, H. and Destexhe, A. (2004) Modeling extracellular field potentials and the frequency-filtering properties of extracellular space. Biophys. J. 64: 1829-1842. [21] Wang K, Riera JJ, Enjieu-Kadji H and Kawashima R. (2013) The role of the extracellular conductivity profiles in compartmental models for neurons: Particulars for layer 5 pyramidal cells. Neural Computation 25: 1807-1852. [22] Poznanski, R. (2010) Thermal noise due to surface-charge effects within the Debye layer of endogenous structures in dendrites. Physical Review E 81: 021902. [23] Conciauro, G. and Puglisi, M. (1981) Meaning of the negative impedance. NASA STI/Recon Technical Report 82: 14458. [24] Mauro, A. (1961) Anomalous impedance, a phenomenological property of time-variant resistance: An analytic review. Biophys. J. 1: 353-372. [25] Braitenberg, V. and A. Shutz. (1998) Cortex: Statistics and Geometry of Neuronal Connectivity. (2nd ed.), Springer-Verlag, Berlin, Germany. [26] Nicholson, C.. (2005) Factors governing diffusing molecular signals in brain extracellular space. J. Neural Transm. 112: 29-44. [27] Hiromu, M, Inoue, M., Miyakawa, H. and Aonishi, T. (2012) Low-Frequency dielectric dispersion of brain tissue due to electri- cally long neurites. Physical Review E 86: 061911. [28] Buzsaki G. (2006) Rhythms of the Brain., Oxford University Press, Oxford UK. [29] Walz W. (1989) Role of glial cells in the regulation of the brain ion microenvironment. Progress Neurobiol. 33: 309-333. [30] Kettenmann H. and Ransom BR. (1995) Neuroglia, Oxford University Press, Oxford, UK. [31] George, M., Lisanby, S., and Sackeim, H. (1999) Transcranial magnetic stimulation : Applications in neuropsychiatry. Arch. Gen. Psychiatry, 56 (4) :300311. [32] Landau, L.D. and ifshitz, E.M. (1984) Electrodynamics of Continuous Media. Pergamon Press, Moscow, Russia. [33] Purcell, E.M. (1985) Electricity and Magnetism. Berkeley Physics Course Vol.2 chap. 3. [34] Note that if the negative charges are exclusively on the membrane (neglecting Debye layers), then the surface integral simplifies to the side of the domain in contact with the membrane. In this case, the electric displacement is different from zero only in the latter portion of the surface. [35] In the presence of magnetic stimulation, such as for example transcranial magnetic stimulation, we have [31 -- 33]: ∇× ~E = − ∂~B , 0 [36] This law can be simplified if one assumes that the cytoplasm is ohmic and homogeneous (zi = ri) for negligible la permittivity. In ∂t this case, we have: i = i f i g i = ~ji f · (πa2 n) = −πa2σe i ∂Vi ∂x 1 ri ∂Vi ∂x = − [37] A medium is said to be "diffusive" when ionic diffusion is non-negligible in the presence of an electric field. [38] For example, if ze = re and zi = ri, then we have R(z(m) e [39] This will be the case for example if two neighboring dendrites have current sources of opposite sign, there will be a direct current flow between them. If they belong to different neurons, this configuration necessarily requires an open-circuit model to be accounted for. ) = − rmre < 0. 1 ri+re 1+ω2τ2 m [40] We use the symbol ≡ for a mathematical identity while the symbol = will mark an equality or a physical law.
1902.06352
1
1902
2019-02-17T23:47:34
Metabolic basis of brain-like electrical signalling in bacterial communities
[ "q-bio.NC", "q-bio.CB" ]
Information processing in the mammalian brain relies on a careful regulation of the membrane potential dynamics of its constituent neurons, which propagates across the neuronal tissue via electrical signalling. We recently reported the existence of electrical signalling in a much simpler organism, the bacterium Bacillus subtilis. In dense bacterial communities known as biofilms, nutrient-deprived B. subtilis cells in the interior of the colony use electrical communication to transmit stress signals to the periphery, which interfere with the growth of peripheral cells and reduce nutrient consumption, thereby relieving stress from the interior. Here we explicitly address the interplay between metabolism and electrophysiology in bacterial biofilms, by introducing a spatially-extended mathematical model that combines the metabolic and electrical components of the phenomenon in a discretised reaction-diffusion scheme. The model is experimentally validated by environmental and genetic perturbations, and confirms that metabolic stress is transmitted through the bacterial population via a potassium wave. Interestingly, this behaviour is reminiscent of cortical spreading depression in the brain, characterised by a wave of electrical activity mediated by potassium diffusion that has been linked to various neurological disorders, calling for future studies on the evolutionary link between the two phenomena.
q-bio.NC
q-bio
Metabolic basis of brain-like electrical signalling in bacterial communities Rosa Martinez-Corral1, Jintao Liu2,*, Arthur Prindle3,*, Gurol M. Suel4, and Jordi Garcia-Ojalvo1 1Department of Experimental and Health Sciences, Universitat Pompeu Fabra, Barcelona Biomedical Research Park, 08003 Barcelona, Spain 2Center for Infectious Diseases Research and Tsinghua-Peking Center for Life Sciences, School of Medicine, Tsinghua University, 100084 Beijing, China 3Department of Biochemistry and Molecular Genetics, Feinberg School of Medicine, Northwestern University, Chicago, IL 60611, USA; Center for Synthetic Biology, Northwestern University, Evanston, 4Division of Biological Sciences, San Diego Center for Systems Biology, and Center for Microbiome Innovation, University of California San Diego, California 92093, USA IL 60208, USA. *Equal contribution Keywords -- membrane potential, electrical signalling, potassium waves, bacterial biofilms, cellular excitability Abstract Information processing in the mammalian brain relies on a careful regulation of the mem- brane potential dynamics of its constituent neurons, which propagates across the neuronal tissue via electrical signalling. We recently reported the existence of electrical signalling in a much simpler organism, the bacterium Bacillus subtilis. In dense bacterial communi- ties known as biofilms, nutrient-deprived B. subtilis cells in the interior of the colony use electrical communication to transmit stress signals to the periphery, which interfere with the growth of peripheral cells and reduce nutrient consumption, thereby relieving stress from the interior. Here we explicitly address the interplay between metabolism and elec- trophysiology in bacterial biofilms, by introducing a spatially-extended mathematical model that combines the metabolic and electrical components of the phenomenon in a discretised reaction-diffusion scheme. The model is experimentally validated by environmental and ge- netic perturbations, and confirms that metabolic stress is transmitted through the bacterial population via a potassium wave. Interestingly, this behaviour is reminiscent of cortical spreading depression in the brain, characterised by a wave of electrical activity mediated by potassium diffusion that has been linked to various neurological disorders, calling for future studies on the evolutionary link between the two phenomena. 1 1 Introduction Local interactions between individual cells are known to generate complex emergent behaviour in multicellular organisms, which underlie the functionalities of tissues, organs, and physiological systems [1]. The human brain provides one of the finest examples of such phenomena, with its complexity ultimately emerging from the interactions between neuronal cells [2]. Similarly, unicellular organisms can also self-organise into communities with complex community-level phenotypes [3]. Bacterial biofilms, in particular, provide a good model system for the study of collective behaviour in biological systems [4]. We have previously shown that biofilms of the bacterium B. subtilis can display community- level oscillatory dynamics [5]. These oscillations ensure the viability, under low-nitrogen (low- glutamate) conditions, of the cells in the centre of the community, which are crucial for com- munity regrowth upon external sources of stress like antibiotics. In our experimental setup (Fig. 1A), growing two-dimensional biofilms that reach a certain size develop oscillations [6], such that peripheral cells periodically stop their growth (see black dotted line in Fig. 1B,C). This behaviour was attributed to a metabolic codependence between central and peripheral cells mediated by ammonium [5] (see introduction to Section 2 below). Besides exhibiting metabolic oscillations in growth rate, our biofilms display periodic changes in extracellular potassium (Fig. 1B) and cellular membrane potential (Fig. 1C), as reported by the fluorescent dyes APG-4 and Thioflavin-T (ThT), respectively. The membrane potential changes reported by ThT are caused by the periodic release of intracellular potassium ions (Ref. [7] and Fig. 1D). Potassium propagates from the centre of the biofilm to its periphery according to the following scenario: potassium released by glutamate-deprived (and thus metabolically stressed) cells in the centre diffuses towards neighbouring cells, causing depolarisation, stress and subsequent potassium release (Fig. 2-right), which thereby keeps propagating along the biofilm in a bucket-brigade manner [7]. Notably, potassium and glutamate are the two most abundant ions in living cells [8]. In addition to their their roles in osmoregulation, pH maintenance and nitrogen metabolism, they are essential for information propagation in animal nervous systems. In particular, potassium is one of the key ions involved in the modulation of the membrane potential in neurons, and glutamate is a central excitatory neurotransmitter. Besides the presence of the same ions in the two scenarios, there are also phenomenological links. In a normally functioning brain, electrical signalling typically propagates along neurons through action potentials mediated by sodium and potassium that can travel at speeds up to meters per second [9]. In addition, under pathological conditions, electrical activity can also propagate over the brain on a much slower timescale, on the order of millimetres per minutes (reviewed in Ref. [10]), similar to the biofilm potassium waves. This occurs during the phenomenon of cortical spreading depression, a wave of intense depolarisation followed by inhibition of electrical activity that entails a strong metabolic disturbance of neuronal function. Cortical spreading depression has been shown to occur in the brains of multiple species, from insects [11, 12] to mammals [13], where it has been regarded as 2 Figure 1: Biofilms of B. subtilis display oscillations in growth rate and electrical signalling activity. A) Scheme of the microfluidics device used to grow biofilms. Biofilms grow around the central pillar or attached to the wall, as in the cartoon. Biofilm dynamics can be represented as a kymograph, where each row represents a one dimensional cross-section of the biofilm area at a given movie frame. B-D) Experimental kymographs of a cross-section of a biofilm, as depicted in A, for different channels. Each pixel is the average of a region of 5.2 × 5.2 µm. The dotted black line denotes the limit of the biofilm as established from the phase-contrast images. B) Oscillations in extracellular potassium, reported by APG-4. C) Oscillations in membrane potential, reported by ThT. D) Extracellular potassium and ThT time traces corresponding to the position denoted by the white line in panels B,C. Data has been smoothed with an overlapping sliding window of size 3 frames. 3 A3 mmspacespacekymograph3 mmspacetime 3 mmspacet=0t=1t=n...spacet=0t=1t=n3 mm3 mmmedia flow0200400600space (m)48121620time (h)Bextracellular potassium0200400600space (m)48121620time (h)Cmembrane potential6101418time (h)0.30.40.50.60.7APG-4 (a.u.)D0.00.20.40.60.81.0APG-4(a.u.)0.00.20.40.60.81.0ThT (a.u.)0.200.250.300.350.400.45ThT (a.u.) an evolutionary conserved process [10] that has been linked to human pathologies like migraine, and to the propagation of brain damage during ischemic stroke. The phenomenon has been modelled as a reaction-diffusion process [14, 15] according to the following mechanism: an initial stimulus leads to an increase in extracellular potassium, causing the membrane potential to rise beyond a critical threshold that triggers a self-amplifying release of potassium, a major ionic redistribution of sodium, calcium and chloride, and release of neurotransmitters like glutamate. Diffusion of potassium and glutamate to neighbouring cells leads to subsequent depolarisation beyond a threshold, thus causing ionic release and self-propagation of the wave [10, 16, 17]. The widespread importance of potassium and glutamate, and the links between electrical signalling in bacteria and neuronal dynamics led us to investigate in more detail the roles of those two ions in the oscillations exhibited by B. subtilis. In this work, we unify in a discretised reaction-diffusion scheme the metabolic and electrical components of the system. Assuming a homogeneous population of cells, the model reveals the spontaneous emergence of two phenotypically-distinct populations of cells as a consequence of the spatiotemporal dynamics of the system, which causes oscillations to start beyond a critical size. We begin by developing a model that accommodates the interactions considered in previous works both for the metabolic and electrical components of the phenomenon. Then we show that the details of the ammonium metabolism are not required to explain the key aspects of the dynamics. The behaviour can be explained by a model with minimal interactions between glutamate metabolism and potassium signalling, thus clarifying the process in bacteria and providing further insight into the roles of these ubiquitous biological ions. 2 A unified spatially-extended model of biofilm oscillations Conceptually, biofilm oscillations can be understood to arise as result of a delayed, spatially- extended negative feedback of metabolic stress [6]. A first molecular model was proposed in Ref. [5] (left half of Fig. 2), based on the fact that cells need to create glutamine from glutamate and ammonium for biomass production. Since ammonium is not present in the MSgg medium used in the experiments, it must be synthesised by the cells from glutamate. The model assumed that only cells in the interior of the biofilm produce ammonium (through the glutamate dehydrogenase enzyme GDH, H in Fig. 2), part of which diffuses away to the periphery of the biofilm, where it is used by the peripheral cells to grow. In turn, peripheral growth reduces glutamate availability in the centre, and subsequently ammonium availability in the periphery. As a result, peripheral cells stop growing, thus allowing the centre to recover glutamate and produce ammonium again, leading to a new oscillation cycle. The model considered two different cell populations -- interior and peripheral -- , and was subsequently implemented on a continuous spatial domain in Ref. [18], with still pre-defined interior and peripheral cell types. This model can readily generate oscillations, which crucially rely on the assumption of a strongly nonlinear activation of GDH by glutamate in the interior population of cells (modelled with a Hill function with large Hill coefficient). However, it does not account for the electrical 4 Figure 2: Metabolic (left) and electric (right) processes involved in the response of the biofilm to a gradient of nutrient-limitation-drive stress from the interior to the periphery. The left-hand side schematises the interactions considered in the metabolic model of Ref. [5], whereas the right-hand side schematises the electrical signalling propagation model proposed in Ref. [7]. In the diagram on the left, H represents the enzyme GDH. In the diagram on the right, EK stands for excess extracellular potassium, V for the membrane potential, and nk for the gating variable of the potassium channel. signalling component of the phenomenon, which we incorporate next. 2.1 Full model description We consider a horizontal cross-section through the middle of a biofilm (as depicted in Fig. 1A) which allows us to simplify the system into a one-dimensional lattice, as shown in Fig. 3. The left-most lattice sites are 'biofilm' (shaded squares, 'b'), followed by 'non-biofilm' sites (white squares, 'n') on the right. We begin by a more detailed model (Fig. 3) that includes all the aforementioned metabolic interactions [5] (left half of Fig. 2) in addition to the electrical signalling propagation [7] (right half of Fig. 2). The details of the equations can be found in the Supplementary, and we next describe the main points of the model (see also Table 1). 'Non-biofilm' sites have extracellular variables only: extracellular glutamate (Ge), ammo- nium (A, assumed to have the same concentration inside and outside cells, following [5]), and extracellular potassium Ke. These variables diffuse between neighboring lattice sites. In ad- dition, we also account for the fact that media is flowing constantly through the microfluidics chamber such that their concentration tends to become close to that in the media (Supplemen- tary Eqs. (S1), (S2), (S17)). 5 30303030interiorperipherygrowthammoniumglutamateammoniumglutamateMetabolic model (Liu et al, 2015)Electric signalling model (Prindle et al, 2015)V stress (S)nkEKEKEKVEKintracellularextracellulardiffusionbiofilmH Figure 3: Scheme of the unified metabolic-electric model introduced in this work. In the lattice scheme on the left, the labels 'b' and 'n' inside the sites denote 'biofilm' and 'non- biofilm' lattice sites, respectively. In the network diagram on the right, G, K, h, H and A represent glutamate, potassium, inactive GDH, active GDH and ammonium, respectively. The quantity r represents biomass-producing biomolecules such as ribosomal proteins. Rg denotes the glutamate receptor, V the membrane potential, nk the gating variable of the potassium channel, and S the stress. h stands for inactive GDH. The subindices 'i' and 'e' label intracellular and extracellular quantities, respectively. The two thick flat arrows highlight the links between the metabolic and electrical components of the system. 'Biofilm' sites correspond to small biofilm regions, such that variable values would corre- spond to averages across multiple cells. As schematised in Fig. 3, in this sites there are also the extracellular variables (also subject to diffusion and media flow) in addition to the intracellular ones that take part in the various biochemical reactions. We model explicitly the dynamics of extracellular (Ge) and intracellular (Gi) glutamate in the biofilm (Supplementary Eqs. (S5)- (S6)), where we consider glutamate uptake into the cells and the conversion of glutamate into ammonium by the action of the enzyme GDH. In turn, we model GDH synthesis in its inactive form (h) and subsequent activation (H) (Supplementary Eqs. (S8)-(S9)). We assume that high concentrations of glutamate inhibit GDH synthesis. This follows from the original observation that GDH overexpression in the periphery kills biofilm oscillations [5], suggesting that GDH expression is restricted to the centre, where glutamate levels are low. Moreover, we assume that there is a threshold for GDH activation by glutamate, as in the original metabolic model. Glutamate and ammonium are assumed to be used for the production of biomass-producing biomolecules such as ribosomal proteins, denoted by r (Supplementary Eq. (S11)). We also ac- count explicitly for the glutamate transporter concentration (Rg), which we assume to saturate for large enough Ge. In order to accommodate the assumption that cells with a higher metabolic activity consume more glutamate, we consider that Rg is subject to an inducible activation that depends on the presence of biomass-producing biomolecules (Supplementary Eq. (S7)). In [7], the YugO potassium channel, which has a TrkA gating domain, was determined to be responsible for potassium release in this phenomenon. TrkA proteins are potassium chan- nel regulatory proteins with nucleotide-binding domains that bind NAD/NADH [19, 20] and ADP/ATP [21], thus coupling potassium translocation to the metabolic state of the cell. Glu- tamate is a central amino acid at the cross-road between carbon and nitrogen metabolism. Our 6 GiRg SnkKiKerAhHVintracellular extracellulardiffusiondiffusionbbbbnnGe assumption is that low glutamate levels lead to imbalances in the aforementioned nucleotides that are sensed by TrkA and trigger potassium release [7]. Therefore, we assume that high levels of intracellular glutamate inhibit the production of stress-related biomolecules, resulting in the following equation for stress dynamics: 1 +(cid:0) Gi S0 (cid:1)ns − γs S (1) dS dt = Gs0 In turn, S enhances the opening probability of the potassium channel, given by nk. This results in a link from the metabolic component of the phenomenon to the electrical counterpart. As the potassium channel opens, potassium is released to the extracellular environment. Conversely, potassium uptake is assumed to be governed by homeostatic processes that tend to keep its intracellular concentration at a fixed value, in addition to depend on the cellular metabolic state (glutamate level), to account for the energy demand of the process (Supplemen- tary Eqs. (S13)-(S14)). Extracellular and intracellular potassium concentrations determine the membrane potential (V ), whose dynamics is described by a Hodgkin-Huxley-like conductance- based model (Supplementary Eq. (S18)). Finally we include the ThT reporter T downstream of the membrane potential, increasing when the cells become hyperpolarised due to potassium release (Supplementary Eq. (S20)). In B. subtilis, glutamate is imported by pumps such as the symporter GltP [22], which uses the proton motive force as a source of energy, and this is in turn influenced by the membrane potential [23]. We therefore consider that glutamate transport into the cell is modulated by the membrane potential V , such that depolarisation reduces entry, and hyperpolarisation enhances it. This provides the link from the electrical to the metabolic component of the phenomenon, and is represented with the following switch-like import-modulation term: F(V ) = 1 1 + exp(gv (V − V0)) (2) The biofilm is assumed to expand proportionally to the biomass-producing biomolecules. In order to simulate growth, we consider that 'non-biofilm' lattice sites neighbouring biofilm sites become occupied with cells (and thus become a lattice site of type 'biofilm') with probability Pgrow dt r, with dt being the simulation time step. Importantly, each new biofilm lattice site inherits the intracellular variables of the 'mother' site. Due to this particularity, the continuum approximation is unsuitable, and we simulate the system as a coupled map lattice. 2.2 Simulations of the model The diffusion coefficient in water for ions and small molecules such as potassium and glutamate is of the order of ∼ 106µm2/h. We thus fixed the diffusion coefficient of potassium and ammonium in the media to this value, and assumed the diffusion coefficient of glutamate to be half that value, due to its larger molecular weight. Regular glutamate concentration in the media is 7 Model assumption Full model (section 2) Simplified model (section 3) Extracellular glutamate flow and diffusion Ammonium flow and diffusion Extracellular potassium flow and diffusion Decay of flow and diffusion rates towards the biofilm interior as a consequence of cellular density Glutamate uptake favoured by hyperpolarisation and disfavoured by depolarisation Yes Yes Yes Yes Yes Yes No (no ammonium present) Yes Yes Yes Glutamate uptake enhanced in metabolically active cells Through the effect of biomass- producing biomolecules Directly linked to intracellular glutamate Saturable glutamate uptake GDH dynamics regulated by glutamate Ammonium produced from intracellular glutamate by GDH Yes Yes Yes Growth favored by high levels of intracellular glutamate Through intermediate biomass- producing biomolecules. Yes No (no GDH present) No (no ammonium present) Directly modulated by intracellular glutamate Metabolic stress inhibited by intracellular glutamate Potassium channel opening probability (nk) increased by stress Potassium released as a result of the opening probability (nk) and membrane potential Potassium uptake governed by homeostatic processes and the metabolic state of the cells (intracellular glutamate) Membrane potential determined by intracellular and extracellular potassium concentrations ThT as a reporter of membrane potential Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Table 1: Main assumptions and comparison between the full and simplified models. 8 30 mM and potassium concentration is 8 mM. Since there is no ammonium in the medium, Am = 0. We fixed the resting membrane potential to −150 mV. Furthermore, oscillations experimentally start at a biofilm size of around 600 µm with a period of around 2 hours [6], and potassium and ThT signals should be correlated [7]. With these constraints, we manually adjusted the rest of the parameters (Table S1) in order to reproduce the experimentally observed oscillatory dynamics. Figure 4 shows kymographs for extracellular potassium and ThT (variables that can be experimentally monitored as explained above, see Fig. 1), as well as for extracellular glutamate, stress, active GDH and ammonium, which have not been experimentally quantified. Oscillations with a period of about two hours emerge in all variables once the biofilm becomes large enough. Moreover, in agreement with the experimental data, the top left plot shows that potassium peaks are likely to coincide with periods of no growth. Figure 4: Oscillations in the full metabolic-electrophysiological model. Kymographs of the model simulation results, showing oscillations in various model variables. In contrast to the original model of Ref. [5] (see also Ref. [18]), where two populations of cells (central and peripheral) were pre-defined regarding GDH production, in the current model an a priori separation of cell types is no longer required, but emerges spontaneously. This can be seen in the bottom left kymograph of Fig. 4, which shows that a central population with higher levels of GDH activity and stress emerges over time. Notably, this model does not require strong cooperativity in GDH activation for the oscillations to emerge (nH = 2, in contrast to the value of 7 in Ref. [5] and 12 in Ref. [18]). This is likely to be the result of the electrical component 9 4812162024time (h)extracellularpotassiumThTextracellular glutamate050010001500space (m)4812162024time (h)active GDH050010001500space (m)ammonium050010001500space (m)stress8910111213141516Ke(mM)0.000.250.500.751.001.251.501.75ThT(µM)5101520Ge(mM)0.000.050.100.150.200.250.300.350.40H(µM)0.30.40.50.60.70.80.9A(mM)0.0000.0050.0100.0150.0200.025S(µM) of the model. In order to test the relevance of this component of the oscillations, we next simplify the metabolic details and consider only the interplay between glutamate metabolism and electrical signalling. 3 Oscillations emerge from the interplay between glutamate and electrical signalling We now simplify the metabolic aspects of the full model introduced above, eliminating GDH, ammonium and biomass-producing biomolecules (Supplementary Eqs. (S8)-(S11)), and suppos- ing instead that intracellular glutamate directly increases the glutamate transporter and directs growth (Fig. 5A). With this, the equations governing the metabolic part of the phenomenon become: dGe dt dGi dt dRg dt = −αg F(V ) Rg = αg F(V ) Rg Ge kg + Ge Ge + Λφ φ (Gm − Ge) + ΛD Dg∇2 Ge − δg Gi Gnr i kg + Ge = αR − δR Rg + αr knr r + Gnr i (3) (4) (5) in addition to considering glutamate diffusion in the 'non-biofilm' sites (Supplementary Eq. S1). In this case, the probability with which the biofilm grows depends directly on the concentra- tion of intracellular glutamate at the biofilm boundary according to Pgrow dt Gi. The dynamics of the electrical part of the model remain unchanged (Supplementary Eqs. (S12)-(S20)). The kymographs in Fig. 5B show that oscillations beyond a critical size also emerge in this case. The results suggest the following mechanism for the oscillations: once the biofilm reaches a critical size, glutamate in the centre of the biofilm is too low and leads to metabolic stress. As a result, cells release potassium, that actively propagates towards the edge due to depolarisation and subsequent metabolic stress. Depolarisation leads to a drop in glutamate consumption levels in the peripheral region, due to the immediate effect on the transport efficiency, and a slightly delayed effect through the reduction in the glutamate transporter. As a result, glutamate can diffuse inwards to allow stress relief in the centre, and a new oscillation cycle can start. According to this, oscillation onset requires stress in the centre to surpass a threshold that leads to potassium release. As we have recently described [6], oscillations can be triggered in experimental biofilms by transiently stopping the media flow of the microfluidics chamber. This is also the case in this model, where oscillations can be triggered by transiently setting the flow-rate parameter (φ) to zero. If the biofilm is sufficiently large, such a perturbation leads to a sudden reduction in the glutamate availability and stress increase, with the subsequent potassium release and oscillatory onset (Fig. S1). Moreover, the model predicts that depolarising any region of the biofilm should be sufficient to trigger a self-propagating wave of stress and potassium release, with the associated changes in membrane potential. This is reproduced experimentally as shown in Fig. 5C: a short increase 10 Figure 5: Oscillations in a simplified model, with glutamate and the electrical signalling compo- nents. A) Scheme of the model. B) Kymographs of various variables of the system. C) Electrical signalling wave (reported by ThT) in an experimental biofilm triggered by an increase of potas- sium in the media from 8 mM to 300 mM during 3 minutes (denoted by the orange mark). The wave propagates inwards, from the periphery (right) to the centre (left). in the concentration of potassium in the media triggers a wave in experimental biofilms, which propagates inwards from the original (peripheral) depolarisation site. 4 Glutamate metabolism and stress release determine oscilla- tion onset size and period According to the aforementioned mechanism for the oscillations, if a biofilm increases its gluta- mate consumption, or if the glutamate concentration in the media is reduced, oscillations should start earlier because the centre becomes stressed at a smaller biofilm size. This expectation is fulfilled when simulations are performed with halved glutamate concentration in the media (Gm), in good agreement with the experimental data (compare the first two conditions in the left panel of Fig. 6, see also Refs. [5, 6]). Similarly, bacteria that cannot synthesise glutamate due to a deletion in the gltA gene are expected to consume more glutamate from the media. We model this condition as an increase in both glutamate uptake rate (αg) and degradation rate (δg), as a result of higher glutamate demand. Also in this case oscillations start earlier both in the model and the experiments (third condition in the left panel of Fig. 6). 11 GiRg SnkKiKeVintracellular extracellulardiffusionbbbnnGeA050100150200space (m)02040time (min)Response to K shock0.00.20.40.60.81.0ThT (a.u.)Cbiofilmnon-biofilm050010001500space (m)4812162024time (h)extracellularpotassium050010001500space (m)membranepotential050010001500space (m)stress81012141618Ke(mM)0.000.250.500.751.001.251.501.75ThT(µM)0.000.010.020.030.040.050.06Stress(µM)B Figure 6: Biofilm size and oscillation period at onset. Light grey: Experimental Data. Dark grey: Model. The data is normalised with respect to the average of each WT condition (model and experiments separately). Boxplots extend between the first and third quartiles of the normalised data, with a line (orange) at the median. The extreme of the whiskers denote the range of the data within 1.5x the interquartile range, with outliers plotted individually. Experimental data contains a minimum of 12 data points per condition. See Methods for details on the simulations. In these two cases of reduced onset size, the onset period is also smaller than in the reference situation (first three conditions in the right panel of Fig. 6). As we have described in previous work [6], larger biofilms tend to have higher periods, the explanation being that the larger a biofilm is, the more time it takes for the stress signal to propagate from centre to periphery, and thus the longer the oscillation period. However, according to the oscillation mechanism that this model suggests, the period of the oscillations must also depend on the metabolism dynamics and the stress relieving capabilites of the cells due to the potassium effects. The ∆trkA mutant strain was characterised in Ref. [7] to be deficient in electrochemical signalling, due to the deletion of the gating domain of the YugO channel. The mutation can be interpreted to render a channel with reduced conductance and increased basal leak [7]. This can be modelled with increased a0 (to simulate a leaky channel), and reduced gK (to simulate reduced conductance once the channel is open). The model simulations predict that oscillations start at the same size as with the basal parameters, but with a reduced period, in good agreement with the experimental data (last condition in Fig. 6). The unaffected onset size is consistent with the fact that glutamate metabolism is largely unperturbed in this mutant, whereas the altered potassium channel leads to less effective stress relief during wave propagation (Figs. S2), and thus facilitates a subsequent cycle. Therefore, the relationship between biofilm size and oscillation period is not universal, but depends upon the metabolic and electrical signalling capabilities of the cells and thus on the genetic background of the strain. 12 WT(1xG)WT(0.5xG)gltAtrkA0.60.81.01.2normalised onset size onset sizeWT(1xG)WT(0.5xG)gltAtrkA0.51.01.5normalised onset period onset periodExperimentModel 5 Discussion The oscillations in growth rate and membrane potential exhibited by B. subtilis biofilms is an instance of self-organising, emergent behaviour similar to the complex collective dynamics that characterizes neuronal tissue. Here we have proposed a unified conceptual framework for these bacterial oscillations that combines glutamate metabolism with potassium wave propagation. Our discrete reaction-diffusion scheme assumes an initially homogeneous population of cells, and exhibits a spontaneous emergence of oscillations beyond a critical biofilm size. Our work shows that oscillations are triggered by a decrease in glutamate levels in the biofilm centre, that lead to metabolic stress and potassium release. Potassium diffusion to the neighbors interferes with glutamate metabolism, thus causing a self-propagating wave of membrane potential changes that allows glutamate recovery in the centre and thus explains the oscillation cycle. With some notable recent exceptions in multicellular eukaryotes [24], electrophysiology has only been considered relevant so far for excitable cells such as those in cardiac and neural tissue. The resemblance, both at the molecular and functional level, between excitable wave propagation in these tissues and the communication of metabolic stress among bacterial cells calls for an evolutionary perspective on the phenomenon of electrical signalling. Animal nervous systems have evolved intricate electrochemical circuits that allow sensing and responding to both internal and external stimuli, with high integrative and computational capabilities clearly evidenced by the human brain. How this complexity arose is still a matter of active investigation [25, 26, 27, 28] but it is tempting to speculate that the bacterial oscillations reported here may represent an ancient instance of electrically-mediated information transmission. Primitive prokaryotic potassium channels are regarded as the ancestors of the animal cation channels whose diversification has been linked to the evolution of nervous systems [29, 30]. Given that high intracellular concentrations of potassium are essential for bacterial pH main- tenance [31, 8], it is natural to expect that potassium channels were used in early bacteria for osmoregulation [27]. The biofilm oscillations studied here suggest that potassium also acts as a signalling intermediate in modern bacteria, communicating information among cells of both the same and also different bacterial species [32]. The link between metabolism and electrical signalling described here is also reminiscent of the relationship between cellular energy and neuronal function [33]. Neuronal dynamics is highly dependent upon cellular metabolism and energy state: the high energy requirements of the ion pumps that maintain proper electrochemical gradients are well known [34]. In addition, some animal voltage-gated ion channels directly respond to energy-related metabolites like NAD(H) and ADP/ATP through direct ligand binding [35], as in the bacterial YugO channel. While the mechanisms that allow information processing in bacterial biofilms share similar chemical and electrical processes with animal brains, the latter compute at much faster speeds and thus in much more complex ways. This can be partly ascribed to the reliance of animal brains on faster voltage-gated ion channels, such as those dependent on sodium, which evolved only when rapid responses were increasingly beneficial (due for instance to the appearance of 13 predators) [27]. Nevertheless, slow propagation of electrical activity is still present in modern animal brains. During the phenomenon of cortical spreading depression, for instance, a slow wave of electrical activity propagates over brain tissue, with massive ionic redistribution involv- ing, critically, potassium and glutamate accompanied by a strong metabolic disturbance. This phenomenon has been regarded as intrinsic to neurons and is evolutionarily conserved across animal species [10]. In the light of these facts, it is tempting to see the potassium-mediated transmission of stress among bacterial cells as a precursor of more recent electrically-mediated information propagation tasks in animal brains. The study of membrane potential dynamics and electrical signalling processes in other bacterial species will be important to further illuminate this issue. Methods 5.1 Experimental data Strains and Plasmids. All experiments were performed using Bacillus subtilis strain NCIB 3610. For the ∆trkA mutant, we deleted the C-terminal portion of yugO (amino acids 117328), leaving only the N-terminal ion channel portion of YugO (amino acids 1116). For the ∆gltA mutant, the gltA gene was replaced by kanamycin resistant gene. Growth conditions. Biofilms were grown using the standard MSgg biofilm-forming medium [36]. In this media glutamate is the only nitrogen source for the bacteria. We explored biofilm dynamics at glutamate concentrations of 1× (30 mM) and 0.5× (15 mM). For microfluidics, We used the CellASIC ONIX Microfluidic Platform and the Y04D microfluidic plate (EMD Millipore). Details can be found in our previous work [5, 37]. Time-Lapse Microscopy. Biofilms were monitored using time-lapse microscopy, using Olym- pus IX81 and IX83 inverted microscopes. To image entire biofilms, 10× lens objectives were used. Images were taken every 10 min. We tracked membrane potential dynamics using the fluorescent dyes Thioflavin T (10 µM) and APG-4 (2µM), from TEFLabs [7]. 5.2 Modeling Simulation methods. Simulations were performed using custom-code in C, and analysis and plotting was done in Python. We used a first-order approximation of the discrete Laplacian operator [38] and the time evolution of the variables was obtained using a 4th order Runge- Kutta method. Boundary conditions were reflective: the outside neighbours of the lattice sites at the boundary of the system are the boundary sites themselves (such that the discrete normal derivatives of the variables at the boundary are zero). The simulation time step was set to 5 × 10−6 h, and each lattice site corresponds to 10 µm. The system was allowed to relax towards the steady state at the beginning of the simulation by preventing expansion during 1 14 hour of simulated time. The initial conditions for the intracellular variables are S = 0 µM , nk = 0, V = −156 mM, Ki = 300 mM, T hT = 0 µM , Rg = 1 µM , Hi = 0.5, H = 0.25, R = 1, and Gi is log-normally distributed around 3 mM. In the biofilm cells, Ke is log-normally distributed around 8 mM, and A = 1 mM. In the non-biofilm cells, A = 0. Initial Ge = 30 mM (or 15 mM in reduced glutamate concentrations). Numerical perturbations. In order to study the effect of halved glutamate in the media, and the trkA and gltA deletions, we assumed the following parameter values with respect to the WT: ∆gltA mutant: δg = 1.5×, αgt = 1.3×; ∆trkA mutant: a0 = 2×, gK = 0.5×. We performed 30 simulations per condition starting from a random initial radius between 150 and 250 µm. Potassium peaks were defined based on a threshold of 9.5 mM (pulse begins when crossing the threshold). Onset period is defined to be the time between the beginnings of the first two peaks, and onset size that of the first beginning. Funding This work was supported by the Spanish Ministry of Economy and Competitiveness and FEDER (project FIS2015-66503-C3-1-P), and by the Generalitat de Catalunya (project 2017 SGR 1054). R.M.C. acknowledges financial support from La Caixa foundation. J.G.O. acknowledges support from the ICREA Academia programme and from the "Mar´ıa de Maeztu" Programme for Units of Excellence in R&D (Spanish Ministry of Economy and Competitiveness, MDM-2014-0370). G.M.S. acknowledges support for this research from the San Diego Center for Systems Biology (NIH grant P50 GM085764), the National Institute of General Medical Sciences (grant R01 GM121888), the Defense Advanced Research Projects Agency (grant HR0011-16-2-0035), and the Howard Hughes Medical Institute-Simons Foundation Faculty Scholars program. A.P. is supported by a Simons Foundation Fellowship of the Helen Hay Whitney Foundation and a Career Award at the Scientific Interface from the Burroughs Wellcome Fund. References [1] J. D. Murray, Mathematical biology. II Spatial models and biomedical applications. Springer, 2001. [2] W. Freeman, Neurodynamics: an exploration in mesoscopic brain dynamics. Springer, 2000. [3] A. F. Mar´ee and P. Hogeweg, "How amoeboids self-organize into a fruiting body: multi- cellular coordination in dictyostelium discoideum," Proceedings of the National Academy of Sciences, vol. 98, no. 7, pp. 3879 -- 3883, 2001. [4] H. C. Flemming, J. Wingender, U. Szewzyk, P. Steinberg, S. A. Rice, and S. Kjelleberg, 15 "Biofilms: An emergent form of bacterial life," Nature Reviews Microbiology, vol. 14, no. 9, pp. 563 -- 575, 2016. [5] J. Liu, A. Prindle, J. Humphries, M. Gabalda-Sagarra, M. Asally, D.-y. D. Lee, S. Ly, J. Garcia-Ojalvo, and G. M. Suel, "Metabolic co-dependence gives rise to collective oscil- lations within biofilms," Nature, vol. 523, no. 7562, pp. 550 -- 554, 2015. [6] R. Martinez-Corral, J. Liu, G. M. Suel, and J. Garcia-Ojalvo, "Bistable emergence of oscillations in growing Bacillus subtilis biofilms," Proceedings of the National Academy of Sciences, vol. 115, pp. E8333 -- E8340, 2018. [7] A. Prindle, J. Liu, M. Asally, S. Ly, J. Garcia-Ojalvo, and G. M. Suel, "Ion channels enable electrical communication in bacterial communities," Nature, vol. 527, no. 7576, pp. 59 -- 63, 2015. [8] J. Gundlach, F. M. Commichau, and J. Stulke, "Perspective of ions and messengers: an in- tricate link between potassium, glutamate, and cyclic di-AMP," Current Genetics, vol. 64, no. 1, pp. 191 -- 195, 2018. [9] J. E. Hall and A. C. Guyton, Guyton and Hall textbook of medical physiology. Elsevier Saunders, 2011. [10] C. Ayata and M. Lauritzen, "Spreading Depression, Spreading Depolarizations, and the Cerebral Vasculature," Physiological Reviews, vol. 95, no. 3, pp. 953 -- 993, 2015. [11] C. I. Rodgers, G. A. B. Armstrong, K. L. Shoemaker, J. D. LaBrie, C. D. Moyes, and R. M. Robertson, "Stress Preconditioning of Spreading Depression in the Locust CNS," PLoS ONE, vol. 2, no. 12, p. e1366, 2007. [12] K. E. Spong, E. C. Rodr´ıguez, and R. M. Robertson, "Spreading depolarization in the brain of Drosophila is induced by inhibition of the Na+/K+-ATPase and mitigated by a decrease in activity of protein kinase G," Journal of Neurophysiology, vol. 116, no. 3, pp. 1152 -- 1160, 2016. [13] M. Lauritzen, J. P. Dreier, M. Fabricius, J. A. Hartings, R. Graf, and A. J. Strong, "Clinical Relevance of Cortical Spreading Depression in Neurological Disorders: Migraine, Malignant Stroke, Subarachnoid and Intracranial Hemorrhage, and Traumatic Brain Injury," Journal of Cerebral Blood Flow & Metabolism, vol. 31, no. 1, pp. 17 -- 35, 2011. [14] M. Dahlem and S. Muller, "Reaction-diffusion waves in neuronal tissue and the window of cortical excitability," Annalen der Physik, vol. 13, no. 78, pp. 442 -- 449, 2004. [15] J. C. Chang, K. C. Brennan, D. He, H. Huang, R. M. Miura, P. L. Wilson, and J. J. Wylie, "A Mathematical Model of the Metabolic and Perfusion Effects on Cortical Spreading Depression," PLoS ONE, vol. 8, no. 8, pp. 1 -- 9, 2013. 16 [16] Y. A. Dahlem, M. A. Dahlem, T. Mair, K. Braun, and S. C. Muller, "Extracellular potas- sium alters frequency and profile of retinal spreading depression waves," Experimental Brain Research, vol. 152, no. 2, pp. 221 -- 228, 2003. [17] R. Enger, W. Tang, G. F. Vindedal, V. Jensen, P. Johannes Helm, R. Sprengel, L. L. Looger, and E. A. Nagelhus, "Dynamics of Ionic Shifts in Cortical Spreading Depression," Cerebral Cortex, vol. 25, pp. 4469 -- 76, 2015. [18] F. Bocci, Y. Suzuki, M. Lu, and J. N. Onuchic, "Role of metabolic spatiotemporal dynamics in regulating biofilm colony expansion," Proceedings of the National Academy of Sciences, vol. 115, no. 16, pp. 4288 -- 4293, 2018. [19] A. Schlosser, A. Hamann, D. Bossemeyer, E. Schneider, and E. P. Bakker, "NAD+binding to the Escherichia coli K+uptake protein TrkA and sequence similarity between TrkA and domains of a family of dehydrogenases suggest a role for NAD+in bacterial transport," Molecular Microbiology, vol. 9, no. 3, pp. 533 -- 543, 1993. [20] T. P. Roosild, S. Miller, I. R. Booth, and S. Choe, "A mechanism of regulating transmem- brane potassium flux through a ligand-mediated conformational switch," Cell, vol. 109, no. 6, pp. 781 -- 791, 2002. [21] Y. Cao, Y. Pan, H. Huang, X. Jin, E. J. Levin, B. Kloss, and M. Zhou, "Gating of the TrkH ion channel by its associated RCK protein TrkA," Nature, vol. 496, no. 7445, pp. 317 -- 322, 2013. [22] F. Kunst et al., "The complete genome sequence of the gram-positive bacterium Bacillus subtilis," Nature, vol. 390, pp. 249 -- 256, 1997. [23] B. Tolner, T. Ubbink-Kok, B. Poolman, and W. N. Konings, "Characterization of the proton/glutamate symport protein of Bacillus subtilis and its functional expression in Es- cherichia coli," Journal of Bacteriology, vol. 177, no. 10, pp. 2863 -- 2869, 1995. [24] M. Levin, G. Pezzulo, and J. M. Finkelstein, "Endogenous bioelectric signaling networks: exploiting voltage gradients for control of growth and form," Annual Review of Biomedical Engineering, vol. 19, pp. 353 -- 387, 2017. [25] L. L. Moroz, "On the independent origins of complex brains and neurons," Brain, Behavior and Evolution, vol. 74, no. 3, pp. 177 -- 190, 2009. [26] D. Bucher and P. A. V. Anderson, "Evolution of the first nervous systems - what can we surmise?," Journal of Experimental Biology, vol. 218, no. 4, pp. 501 -- 503, 2015. [27] W. B. Kristan, "Early evolution of neurons," Current Biology, vol. 26, no. 20, pp. R949 -- R954, 2016. 17 [28] L. L. Moroz and A. B. Kohn, "Independent origins of neurons and synapses: insights from ctenophores," Philosophical Transactions of the Royal Society B: Biological Sciences, vol. 371, no. 1685, p. 20150041, 2016. [29] P. A. V. Anderson and R. M. Greenberg, "Phylogeny of ion channels: Clues to structure and function," Comparative Biochemistry and Physiology - B Biochemistry and Molecular Biology, vol. 129, no. 1, pp. 17 -- 28, 2001. [30] H. H. Zakon, "Adaptive evolution of voltage-gated sodium channels: The first 800 million years," Proceedings of the National Academy of Sciences, vol. 109, pp. 10619 -- 10625, 2012. [31] C. M. Armstrong, "Packaging Life: The Origin of Ion-Selective Channels," Biophysical Journal, vol. 109, no. 2, pp. 173 -- 177, 2015. [32] J. Humphries, L. Xiong, J. Liu, A. Prindle, F. Yuan, H. A. Arjes, L. Tsimring, and G. M. Suel, "Species-Independent Attraction to Biofilms through Electrical Signaling," Cell, vol. 168, no. 1-2, pp. 200 -- 209.e12, 2017. [33] L. Yu and Y. Yu, "Energy-efficient neural information processing in individual neurons and neuronal networks," Journal of Neuroscience Research, vol. 95, no. 11, pp. 2253 -- 2266, 2017. [34] P. J. Magistretti and I. Allaman, "A Cellular Perspective on Brain Energy Metabolism and Functional Imaging," Neuron, vol. 86, no. 4, pp. 883 -- 901, 2015. [35] P. J. Kilfoil, S. M. Tipparaju, O. A. Barski, and A. Bhatnagar, "Regulation of ion channels by pyridine nucleotides," Circulation Research, vol. 112, no. 4, pp. 721 -- 741, 2013. [36] S. S. Branda, J. E. Gonz´alez-Pastor, S. Ben-Yehuda, R. Losick, and R. Kolter, "Fruiting body formation by bacillus subtilis," Proceedings of the National Academy of Sciences, vol. 98, no. 20, pp. 11621 -- 11626, 2001. [37] J. Liu, R. Martinez-Corral, A. Prindle, D. L. Dong-yeon, J. Larkin, M. Gabalda-Sagarra, J. Garcia-Ojalvo, and G. M. Suel, "Coupling between distant biofilms and emergence of nutrient time-sharing," Science, vol. 356, pp. 638 -- 642, 2017. [38] F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, and B. V. Saunders, "NIST Digital Library of Mathematical Func- tions." Release 1.0.18, 2018. 18 SUPPLEMENTARY MATERIAL 1 Model equations and description As explained in the main text, we consider a one-dimensional array of simulation lattice sites where biochemical species react and diffuse. Here we describe in detail the model equations. 1.1 Metabolic component In the 'non-biofilm' sites, we consider media flow and diffusion of glutamate (G) and ammonium (A) according to the following equations: dGe dt dA dt = φ (Gm − Ge) + Dg∇2 Ge = φ (Am − A) + Da∇2 A (S1) (S2) In the microfluidics chamber, media is flowing constantly. Thus, we simulate the effect of the media flow with the first term of each equation, such that with some rate φ, the concentration of the chemical species tends to equate that in the medium (Xm). The second term models diffusion. In the biofilm, we assume that the diffusion coefficient and flow rate of glutamate decay exponentially with the distance to the biofilm edge de, due to the extracellular matrix and high cell density, according to the following functions: Λφ = ΛD = exp(−γφ de) + aφ 1 + aφ exp(−γD de) + aD 1 + aD (S3) (S4) These functional forms are chosen such that at the edge Λx = 1, and flow and diffusion match those in the media. The coefficients decay towards the biofilm interior, tending asymptotically to a/(1 + a) in the centre, which ensures some remaining flow and diffusion. We follow the original assumption [5] that ammonium diffusion over the biofilm is very fast, and do not apply these reduction terms to this variable. The dynamics of extracellular (Ge) and intracellular (Gi) glutamate in the biofilm are mod- elled with the following dynamical equations: dGe dt dGi dt = −αg F(V ) Rg = αg F(V ) Rg kg + Ge Ge kg + Ge Ge + Λφ φ (Gm − Ge) + ΛD Dg∇2 Ge − αa H Gi − δg Gi r , (S5) (S6) where, as mentioned, Ge is affected by media flow and diffusion. The first term in the right-hand side of the two equations represents glutamate transport into the cells. As explained in the main text, we consider that glutamate transport into the cell is modulated by the membrane potential V , such that depolarisation reduces entry, and hyperpolarisation enhances it, according to the functional form given in Eq. (2) of the main text. 19 In addition, we assume that glutamate is imported into the cells through the glutamate transporter Rg, which saturates for large enough Ge, with half-maximum concentration kg. We describe explicitly the dynamics of Rg by: dRg dt = αR − δR Rg + βR rnR knR R + rnR . (S7) We thus assume that Rg is produced at a basal rate αR and degraded at a rate δR. The last term accounts for the higher glutamate uptake by metabolically active cells, such that the presence of biomass-producing biomolecules, such as ribosomal proteins, denoted by r, enhances Rg synthesis via a Hill function with exponent nR. Equation (S6) also assumes that intracellular glutamate concentration decays due to ammo- nium production via the GDH enzyme (represented by H in the αa-term at the right-hand side of the equation) and through various metabolic tasks including in particular biomass production (δg-term). The production of ammonium is regulated by the activity of the enzyme GDH. We describe the dynamics of the inactive and active forms of this enzyme, h and H respectively, by the equations: dh dt dH dt αh = 1 + (Gi/kh)nh GnH i = αH h H + GnH knH i − αH h GnH i knH H + GnH − γH H i − γh h + γH H (S8) (S9) such that we account for synthesis and degradation of inactive GDH and its conversion into the active form (αH -term in the two equations). As explained in the main text, we assume that high concentrations of glutamate inhibit GDH synthesis, whereas activation is positively regulated by glutamate via a Hill function with exponent nH . We also consider deactivation at a constant rate γH . Ammonium dynamics is affected by production from glutamate by active GDH, consumption for various metabolic processes such as biomass production, and diffusion: = αa H Gi − δa A r + Da∇2 A dA dt (S10) Finally, biomass production is considered to increase with ammonium and intracellular gluta- mate, and to be subject to linear decay: = βr A Gi − γr r dr dt (S11) 1.2 Electrical signalling Next we incorporate an adapted version of the electrical model introduced in [7]. As explained in the main text, we consider an inhibitory effect of intracellular glutamate on a stress variable S, whose production rate is modelled with an inhibitory Hill function: 1 +(cid:0) Gi S0 Gs0 (cid:1)ns − γs S dS dt = (S12) 20 We explicitly consider both extracellular (Ke) and intracellular (Ki) potassium, whose dy- namics are given by: dKe dt dKi dt = F gK n4 = −F gK n4 k (V − VK) − Dp Gi Ke(Ki0 − Ki) + Λφ φ (Km − Ke) + ΛDDk∇2 Ke k (V − VK) + Dp Gi Ke(Ki0 − Ki) (S13) (S14) Potassium uptake is assumed to be governed by homeostatic processes that tend to keep its intracellular concentration at a fixed value Ki0, described by the second term in the right-hand side of Eqs. (S13)-(S14). Uptake is also made to depend on the metabolic state (glutamate level), to account for the energy demand of the process. In addition, extracellular potassium diffuses and is subject to the media flow in the chamber. Potassium flow through its ion channel (first term in the right-hand side of the Ke and Ki equations) is governed by the corresponding Nernst potential: (cid:18) Ke (cid:19) VK = VK0 ln , Ki (S15) (S16) and depends on the opening probability of the potassium channel, nk: dnk dt = a0 S Sth + S (1 − nk) − b nk In the media lattice sites, as in the case of glutamate and ammonium [Eqs. (S1) and (S2)], the extracellular potassium dynamics is affected by diffusion and by the media flow: dKe dt = φ (Km − Ke) + Dk∇2 Ke (S17) The membrane potential dynamics is described by a Hodgkin-Huxley-like conductance-based model containing potassium flux through the ion channel and a leak current [7]: dV dt = −gK n4 k (V − VK) − gL (V − VL) , (S18) where the leak potential VL is assumed to depend on the extracellular potassium [7] in a threshold-linear manner, such that when Ke is larger than its basal level in the medium, Km, the leak potential VL grows linearly (and the cell depolarizes), while when Ke < Km the leak potential stays at its basal level: VL = VL0 + dL Ke − Km 1 − e−(Ke−Km)/σ (S19) Finally, we include the ThT reporter T downstream of the membrane potential, increasing when the cells become hyperpolarised due to potassium release: dT dt = αT 1 + exp(gT (V − V0T )) − γT T (S20) 21 1.3 Simplified model In the simplified model, we keep the same dynamics for the electrical part (Eqs. (S12)-(S20)). The metabolic part is simplified as follows: Eqs. (S8)-(S11) are removed, the equations for extracellular glutamate dynamics, in both 'biofilm' and 'non-biofilm' sites (Eqs. S1, S5), are maintained, and Eqs. (S6) and (S7) become: dGi dt dRg dt = αg F(V ) Rg Ge kg + Ge = αR − δR Rg + αr − δg Gi Gnr i knr r + Gnr i (S21) (S22) 2 Supplementary tables and figures Parameter Description Full model Simplified Units αg kg Gm Dg αa δg αR δR βR kR nR αh kh γh nh αH kH nH γH δa Am Da glutamate uptake constant extracellular glutamate concentration at half- maximal uptake rate glutamate concentration in the media glutamate diffusion coefficient ammonium production constant glutamate degradation constant glutamate receptor synthesis rate glutamate receptor decay constant maximum rate of glutamate receptor induc- tion threshold for Rg induction Hill coefficient for Rg induction maximal GDH synthesis rate 36.0 0.75 30.0 4e+06 4.5 0.525 6.75 36.0 45.0 5.0 2.0 0.075 threshold for inhibition of GDH synthesis by 1.5 glutamate GDH decay rate 0.01 Hill coefficient for GDH synthesis inhibition 2.0 by glutamate GDH activation constant intracellular glutamate concentration for half- maximal GDH activation 3.0 0.4 Hill coefficient for GDH activation by gluta- 2.0 mate GDH deactivation constant ammonium consumption constant ammonium concentration in the media ammonium diffusion coefficient 5.0 0.135 0.0 7e+06 model 24.0 0.75 30.0 4e+06 - 4.8 4.5 24.0 31.0 2.25 2.0 - - - - - - - - - - - mM /(µM h) mM mM µm2/h µM−1h−1 mM−1h−1 µM/h h−1 µM/h mM - µM/h mM h−1 - h−1 mM - h−1 mM−1h−1 mM µm2/h Table S1 -- Continued on next page 22 Parameter Description Full model Simplified Units Table S1 -- Continued from previous page βr γr S0 GS0 ns γs a0 Sth b F gK VK0 Dp Ki0 Dk Km gL dL VL0 σ gv V0 αT γT gT Pgrow γφ aφ γD aD φ model - - 1.12 0.2 2.0 2.8 91.0 0.03 34.0 0.05 70.0 100.0 0.12 - 7e+06 8.0 18.0 4.0 mM−1h−1 h−1 µM/h mM - h−1 h−1 µM h−1 mM/mV h−1 mV mM−2h−1 mM µm2/h mM h−1 mV/mM mV mM mV−1 mV µM/h h−1 mV−1 h−1 µm - µm h−1 biomass-producing biomolecules synthesis 15.0 rate biomass-producing biomolecules decay rate maximum stress production rate threshold for stress inhibition by glutamate Hill coefficient for stress inhibition by gluta- mate stress decay constant 6.0 1.12 0.2 2.0 2.8 maximum rate of increase of the potassium 91.0 channel gating probability stress level for half-maximal gating activity opening probability decay constant membrane capacitance potassium channel conductance Nernst potential prefactor potassium uptake constant intracellular potassium concentration potassium diffusion coefficient potassium concentration in the media leak conductance leak slope coefficient basal leak potential 0.03 21.25 0.05 70.0 100.0 0.12 300.0 7e+06 8.0 18.0 4.0 -156.0 -156.0 leak threshold sharpness coefficient inverse sensitivity of glutamate uptake to membrane potential resting membrane potential maximal rate of ThT uptake intracellular ThT decay constant 0.1 1.0 -150.0 20.0 10.0 inverse sensitivity of ThT to membrane poten- 0.3 tial growth probability 0.3 spatial decay rate of the flow rate within the 0.0085 biofilm basal flow factor 0.012 spatial decay rate of the diffusion coefficient 0.0085 within the biofilm basal diffusion factor flow rate 0.012 5.0 0.1 1.0 -150.0 20.0 10.0 0.3 0.5 0.0085 0.012 0.0085 - 5.0 Table S1: Parameter description and basal values for the two mod- els. 23 Figure S1: Stop-flow triggers oscillations in the simplified model. Normalised ThT time traces at the periphery (50 µm from the biofilm edge). A) Reference simulation. Vertical dashed line indicates the time of stop-flow in B. (This simulation is the same as in Fig. 5B). B) The biofilm was perturbed with a stop-flow-like perturbation (φ = 0 during 20 min at t = 9 hours). The growth noise realisation is the same in both simulations, such that the only difference is the perturbation. Figure S2: The ∆trkA mutation in the model leads to impaired stress relief. For each simulation, either the maximum or the minimum of the variable during the oscillations was computed for the peripheral region (outermost 100 µm). Each dot represents a simulation, from the same data as in Fig. 6 from the main text. 24 5101520time (h)0.00.51.0ThT (a.u.)5101520time (h)0.00.51.0ThT (a.u.)ABWTtrkA0.0500.0750.1000.125minnkWTtrkA15.017.520.022.5maxKe(mM)WTtrkA0.00060.0008minS(mM)
1111.3610
1
1111
2011-11-15T19:01:36
Scaling of brain metabolism and blood flow in relation to capillary and neural scaling
[ "q-bio.NC", "physics.bio-ph", "q-bio.CB", "q-bio.TO" ]
Brain is one of the most energy demanding organs in mammals, and its total metabolic rate scales with brain volume raised to a power of around 5/6. This value is significantly higher than the more common exponent 3/4 relating whole body resting metabolism with body mass and several other physiological variables in animals and plants. This article investigates the reasons for brain allometric distinction on a level of its microvessels. Based on collected empirical data it is found that regional cerebral blood flow CBF across gray matter scales with cortical volume $V$ as $CBF \sim V^{-1/6}$, brain capillary diameter increases as $V^{1/12}$, and density of capillary length decreases as $V^{-1/6}$. It is predicted that velocity of capillary blood is almost invariant ($\sim V^{\epsilon}$), capillary transit time scales as $V^{1/6}$, capillary length increases as $V^{1/6+\epsilon}$, and capillary number as $V^{2/3-\epsilon}$, where $\epsilon$ is typically a small correction for medium and large brains, due to blood viscosity dependence on capillary radius. It is shown that the amount of capillary length and blood flow per cortical neuron are essentially conserved across mammals. These results indicate that geometry and dynamics of global neuro-vascular coupling have a proportionate character. Moreover, cerebral metabolic, hemodynamic, and microvascular variables scale with allometric exponents that are simple multiples of 1/6, rather than 1/4, which suggests that brain metabolism is more similar to the metabolism of aerobic than resting body. Relation of these findings to brain functional imaging studies involving the link between cerebral metabolism and blood flow is also discussed.
q-bio.NC
q-bio
Scaling of brain metabolism and blood flow in relation to capillary and neural scaling 1 1 0 2 v o N 5 1 ] . C N o i b - q [ 1 v 0 1 6 3 . 1 1 1 1 : v i X r a Jan Karbowski Institute of Biocybernetics and Biomedical Engineering, Polish Academy of Sciences, 02-109 Warsaw, Poland Abstract Brain is one of the most energy demanding organs in mammals, and its total metabolic rate scales with brain volume raised to a power of around 5/6. This value is significantly higher than the more com- mon exponent 3/4 relating whole body resting metabolism with body mass and several other physiological variables in animals and plants. This article investigates the reasons for brain allometric distinction on a level of its microvessels. Based on collected empirical data it is found that regional cerebral blood flow CBF across gray matter scales with cortical volume V as CBF ∼ V −1/6, brain capillary diam- eter increases as V 1/12, and density of capillary length decreases as V −1/6. It is predicted that velocity of capillary blood is almost in- variant (∼ V ǫ), capillary transit time scales as V 1/6, capillary length increases as V 1/6+ǫ, and capillary number as V 2/3−ǫ, where ǫ is typ- ically a small correction for medium and large brains, due to blood viscosity dependence on capillary radius. It is shown that the amount of capillary length and blood flow per cortical neuron are essentially conserved across mammals. These results indicate that geometry and dynamics of global neuro-vascular coupling have a proportionate char- acter. Moreover, cerebral metabolic, hemodynamic, and microvascular variables scale with allometric exponents that are simple multiples of 1/6, rather than 1/4, which suggests that brain metabolism is more similar to the metabolism of aerobic than resting body. Relation of these findings to brain functional imaging studies involving the link between cerebral metabolism and blood flow is also discussed. 1 Keywords: Brain metabolism; Cerebral blood flow; Capillary; Scaling; Mammals. Email: [email protected]; [email protected] 2 Introduction It is well established empirically that whole body metabolism of resting mammals scales with body volume (or mass) with an exponent close to 3/4, which is known as Kleiber's law [1, 2, 3, 4]. The same exponent or its simple derivatives govern the scalings of respiratory and cardiovascular systems in mammals and some other physiological parameters in animals and plants [2, 3, 5]. Because of its almost ubiquitous presence, the quarter power has often been described as a general law governing metabolism and blood circulation, and several formal models explaining its origin have been proposed that still cause controversy [6, 7, 8, 9]. However, as was found by the author [10], the brain metabolism at rest seems to follow another scaling rule. Total brain metabolic rate (both oxygen and glucose) scales with brain volume with an exponent ≈ 0.85, or close to 5/6 [10]. Consequently, the volume-specific cerebral metabolism decreases with brain size with an exponent around −1/6, and this value is highly homogeneous across many structures of gray matter [10]. The origin of these cerebral exponents has never been explained, although it is interesting why brain metabolism scales different than metabolism of other systems. The brain, similar to other organs, uses capillaries for delivery of metabolic nutri- ents (oxygen, glucose, etc.) to its cells [11]. Moreover, numerical density of cerebral capillaries is strongly correlated with brain hemodynamics and metabolism [12, 13]. However, the cerebral microvascular network differs from other non-cerebral networks in two important ways. First, in the brain there exists a unique physical border, called the brain-blood barrier, which severely restricts influx of undesired molecules and ions 3 to the brain tissue. Second, cerebral capillaries exhibit a large degree of physical plastic- ity, manifested in easy adaptation to abnormal physiological conditions. For instance, during ischemia (insufficient amount of oxygen in the brain) capillaries can substantially modify their diameter to increase blood flow and hence oxygen influx [14, 15, 16]. These two factors, i.e. structural differences and plasticity of microvessels, can in principle modify brain metabolism in such a way to yield different scaling rules in comparison to e.g. lungs or muscles. Another, related factor that may account for the uncommon brain metabolic scaling is the fact that brain is one of the most energy expensive organs in the body [10, 17]. This is usually attributed to the neurons with their extended axons and dendrites, which utilize relatively large amounts of glucose and ATP for synaptic communication [18, 19]. The main purpose of this paper is to determine scaling laws for blood flow and geometry of capillaries in the brain of mammals. Are they different from those found or predicted for cardiovascular and respiratory systems? If so, do these differences account for brain metabolic allometry? How the scalings of blood flow and capillary dimensions relate to the scalings of neural characteristics, such as neural density and axon (or dendrite) length? This study might have implications for expanding of our understanding of mammalian brain evolution, in particular the relationship between brain wiring, metabolism, and its underlying microvasculature [10, 20, 21]. The results can also be relevant for research involving the microvascular basis of brain functional imaging studies, which use relationships between blood flow and metabolism to decipher regional neural activities [22, 23]. 4 Results The data for brain circulatory system were collected from different sources (see Materials and Methods). They cover several mammals spanning 3-4 orders of magnitude in brain volume, from mouse to human. 1. Empirical scaling data. Cerebral blood flow CBF in different parts of mammalian gray matter decreases systematically with gray matter volume, both in the cortical and subcortical regions (Fig. 1). In the cerebral cortex, the scaling exponent for regional CBF varies from −0.13 for the visual cortex (Fig. 1A), −0.15 for the parietal cortex (Fig. 1B), −0.17 for the frontal cortex (Fig. 1C), to −0.19 for the temporal cortex (Fig. 1D). The average cortical exponent is −0.16 ± 0.02. In the subcortical regions, the CBF scaling exponent is −0.14 for hippocampus (Fig. 2A), −0.17 for thalamus (Fig. 2B), and −0.18 for cerebellum (Fig. 2C). The average subcortical exponent is identical with the cortical one, i.e., −0.16 ± 0.02, and both of them are close to −1/6. It is interesting to note that almost all of the cortical areas (except temporal cortex) have scaling exponents whose 95% confidence intervals do not include a quarter power exponent −1/4. The microvessel system delivering energy to the brain consists of capillaries. The capillary diameter increases very weakly but significantly with brain size, with an ex- ponent of 0.08 (Fig. 3A). On the contrary, the volume-density of capillary length decreases with brain size raised to a power of −0.16 (Fig. 3B). Thus, the cerebral capillary network becomes sparser as brain size increases. Despite this, the fraction of gray matter volume taken by capillaries is approximately independent of brain size 5 (Fig. 3C). Another vascular characteristic, the arterial partial oxygen pressure, is also roughly invariant with respect to brain volume (Fig. 3D). A degree of neurovascular coupling can be characterized by geometric relationships between densities of capillaries and neurons. Scaling of the density of neuron number in the cortical gray matter is not uniform across mammals [24, 25, 26]. In fact, the scaling exponent depends to some extent on mammalian order and the animal sample used [26]. For the sample of mammals used in this study, it is found that cortical neuron density decreases with cortical gray matter volume with an exponent of −0.13 (Fig. 4A). This exponent is close to the exponent for the scaling of capillary length density, which is −0.16 (Fig. 3B). Consistent with that, the ratio of cortical capillary length density to neuron density across mammals is approximately constant and independent of brain size (Fig. 4B). Typically, there is about 10 µm of capillaries per cortical neuron. The scaling dependence between the two densities yields an exponent close to unity (Fig. 4C), which shows a proportionality relation between them. Cerebral blood flow CBF scales with brain volume the same way as does capillary length density (Figs. 1,2,3B), and thus, CBF should also be related to neural density. Indeed, in the cerebral cortex the ratio of the average CBF to cortical neural density is independent of brain scale (Fig. 5). This means that the average amount of cortical blood flow per neuron is invariant among mammals, and about (1.45±0.4)10−8 mL/min. Taken together, the findings in Figs. 4 and 5 suggest a tight global correlation between neurons and their energy supporting microvascular network. 6 2. Theoretical scaling rules for cerebral capillaries. Below I derive theoretical predictions for the allometry of brain capillary character- istics, such as: capillary length and radius, capillary number, blood velocity, and time taken by blood to travel through a capillary. I also find relationships connecting cere- bral metabolic rate and blood flow with neuron density. The following assumptions are made in the analysis: (i) Oxygen consumption rate in gray matter CMRO2 scales with cortical gray matter volume V as V −1/6, in accordance with Ref. [10]; (ii) Capillary volume fraction, fc = πNcLcR2 c/V , is invariant with respect to V , which follows from the empirical results in Fig. 3C. The symbol Nc denotes total capillary number in the gray matter, Lc is the length of a single capillary segment, and Rc is its radius; (iii) Driving blood pressure ∆pc through capillaries is independent of brain size, which is consistent with a known fact that arterial blood pressure (both systolic and diastolic) of resting mammals is independent of body size [27, 28, 29]; (iv) Partial oxygen pressure pO2 in capillaries is also invariant, which is consistent with the empirical data in Fig. 3D on the invariance of arterial oxygen pressure; (v) Cerebral blood flow CBF is pro- portional to oxygen consumption rate CMRO2, due to adaptation of capillary diameters to oxygen demand. The cerebral metabolic rate of oxygen consumption CMRO2, according to the modi- fied Krogh model [11, 14], is proportional to the product of oxygen flux through capillary wall and the tissue-capillary gradient of oxygen pressure ∆pO2, i.e. CMRO2 ∼ D(cid:18)NcLc V (cid:19) ∆pO2, (1) 7 where D is the oxygen diffusion constant in the brain. The dependence of CMRO2 on capillary radius in this model has mainly a logarithmic character, and hence it is neglected as weak. Since oxygen pressure in the brain tissue is very low [30], the pressure gradient ∆pO2 is essentially equal to the capillary oxygen pressure pO2. Consequently, the formula for CMRO2 simplifies to CMRO2 ∼ ρcpO2, where ρc is the density of capillary length ρc = NcLc/V . From the assumptions (i) and (iv) we obtain that capillary length density ρc ∼ V −1/6. Additionally, from (ii) we have R2 c ∼ fc/ρc ∼ V 0/V −1/6 ∼ V 1/6, implying that capillary radius (or diameter) Rc scales as V 1/12. Consequently capillary diameter does not increase much with brain magnitude. As an example, a predicted capillary diameter for elephant with its cortical gray matter volume 1379 cm3 [31] is 7.2 µm, which does not differ much from those of rat (4.1 µm [15, 32]) or human (6.4 µm [33, 34]), who have corresponding volumes 3450 and 2.4 times smaller. The blood flow Qc through a capillary is governed by a modified Poiseuille's law in which blood viscosity depends on capillary radius [35]: Qc = π∆pcR4 c 8ηef (Rc)Lc , (2) where ∆pc is the axial driving blood pressure along a capillary of length Lc, and ηef (Rc) is the capillary radius dependent effective blood viscosity. The latter dependence has a nonmonotonic character, i.e. for small diameters the viscosity ηef (Rc) initially decreases 8 with increasing Rc, reaching a minimum at diameters about 5 − 7 µm. For 2Rc > 10 µm the blood viscosity ηef slowly increases with Rc approaching its bulk value for diameters ∼ 500 µm. This phenomenon is known as the Fahraeus-Lindqvist effect [36]. In general, blood viscosity in narrow microvessels depends on microvessel thickness because red blood cells tend to deform and place near the center of capillary leaving a cell-free layer near the wall [35, 37]. These two regions have significantly different viscosities, with the cell-free layer having essentially plasma viscosity ηp, which is much smaller than the bulk (or center region) viscosity ηc. The formula relating the effective blood viscosity ηef with capillary radius and both viscosities ηp and ηc is given by [35]: ηef (Rc) = ηp 1 − (1 − µ)(1 − w/Rc)4 , (3) where µ = ηp/ηc ≪ 1, and w is the thickness of cell-free layer. For capillary radiuses relevant for the brain, i.e. 1.5 µm < Rc < 3.5 µm (see Suppl. Table S2), the ratio w/Rc increases with increasing Rc, which causes a decline in the effective blood viscosity down to its minimal value at Rc = 3 − 3.5 µm (Table 1). Using the data in Table 1 taken from [35], we can approximate the denominator in Eq. (3) for this range of radiuses by a simple, explicit function of Rc. The best fit is achieved with a logarithmic function, i.e. 1 − (1 − µ)(1 − w/Rc)4 ≈ 0.85[ln(Rc/Ro)]2/3, where Ro = 1.2 µm (Table 1). As a result, the effective blood viscosity takes a simple form: ηef (Rc) = 1.18ηp (ln(Rc/Ro))−2/3 . (4) 9 Cerebral blood flow CBF in the brain gray matter is defined as CBF = Q/V , where Q = NcQc is the total capillary blood flow through all Nc capillaries. Thus CBF is given by or CBF ∼ ∆pc ηef (Rc) ρcR4 c L2 c , CBF ∼ ∆pcρcR4 c ηpL2 c (cid:20)ln(cid:18) Rc Ro(cid:19)(cid:21) 2/3 . (5) (6) We can rewrite the logarithm present in Eq. (6), in an equivalent form, as a power function (Rc/Ro)γ with a variable exponent γ given by (see Appendix S1 in the Supp. Infor.): so that CBF becomes γ = 2 3 ln(ln(Rc/Ro)) ln(Rc/Ro) , CBF ∼ ∆pc ηp ρcR4+γ c L2 c . (7) (8) The exponent γ in this equation can be viewed as a correction due to non-constant 10 blood viscosity (Fahraeus-Lindqvist effect [36]). The dependence of γ on the capillary diameter is shown in Table 2. Because in general γ < 0, its presence in Eq. (8) reduces the power of Rc. However, this effect is weak for medium and large brains as γ ≪ 1. Even for a small rat brain the relative influence of γ is rather weak, since γ/4 ≈ 0.19. In contrast, for very small brains, such as mouse, the effect caused by γ is strong (Table 2), which reflects a sharp increase in the effective blood viscosity for the smallest capillaries [35, 37]. Now we are in a position to derive scaling rules for the capillary length segment Lc, capillary blood velocity uc, and the number of capillaries Nc. From Eq. (8), using the assumptions (i), (iii), and (v), we obtain L2 c ∼ ρcR4+γ c /CMRO2, which implies that Lc ∼ R2+γ/2 c (viscosity of blood plasma is presumably independent of brain scale [38]). Consequently, Lc ∼ V 1/6+γ/24, i.e. capillary length should weakly increase with brain size. Although there are no reliable data on Lc, we can compare our prediction with the measured intercapillary distances, which generally should be positively correlated with Lc. Indeed, the mean intercapillary distance in gray matter increases with increasing brain volume, and is 17 − 24 µm in rat [39], 24 µm in cat [40], and 58 µm in human [33]. Average velocity uc of blood flow in brain capillaries is given by uc = Qc/(πR2 c). Using the expressions for Qc and ηef , we get uc ∼ (∆pcR2+γ c )/Lc. Since Lc ∼ R2+γ/2 c above, and using the assumption (iii), we obtain uc ∼ Rγ/2 c ∼ V γ/24. Thus, capillary blood velocity is almost independent of brain size for medium and large brains, as then γ → 0 (Table 2). For very small brains, instead, there might be a weak dependence. A 11 related quantity, the blood transit time τc through a capillary, defined as τc = Lc/uc, scales as τc ∼ V 1/6, regardless of the brain magnitude. This indicates that τc and CBF are inversely related across different species, τc ∼ CBF−1, because of their scaling properties. We can find the scaling relation for the total number of capillaries Nc from the volume-density of capillary length ρc. We obtain Nc = ρcV /Lc ∼ V −1/6V /V 1/6+γ/24 ∼ V 2/3−γ/24, i.e. the exponent for Nc is close to 2/3 for not too small brains. As an example, the number of capillary segments in the human cortical gray matter should be 123 times greater than that in the rat (cortical volumes of both hemispheres in rat and human are 0.42 cm3 [24] and 572.0 cm3 [41], respectively). As was shown above, CMRO2 must be proportional to the volume density of capillary length ρc (Eq. 1). On the other hand, the empirical results in Fig. 4 indicate that ρc is roughly proportional to neuron density ρn. Thus, we have approximately CMRO2 ∼ ρn across different mammals. This implies that oxygen metabolic energy per neuron in the gray matter should be approximately independent of brain size. Exactly the same conclusion was reached before in a study by Herculano-Houzel [26], based on independent data analysis. Moreover, since cortical CMRO2 and CBF scale the same way against brain size, we also have CBF ∼ ρn, which is confirmed by the results in Fig. 5. In other words, both cerebral metabolic rate and blood flow per neuron are scale invariant. 12 Discussion 1. General discussion. The summary of the scaling results is presented in Table 3. Some of these allomet- ric relations are directly derived from the experimental data (CBF, Rc, ρc, fc, ρc/ρn, CBF/ρn), and others are theoretically deduced (Nc, Lc, uc, τc). The interesting result is that cerebral blood flow CBF in gray matter scales with cortical gray matter volume raised to a power of −0.16. The similar exponent governs the allometry of cortical metabolic rate CMR [10], which indicates that brain metabolism and blood flow are roughly linearly proportional across different mammals. This conclusion is compatible with several published studies that have shown the proportionality of CMR and CBF on a level of a single animal (rat, human) across different brain regions [12, 42]. The coupling between CMR and CBF manifests itself also in their relation to the number of neurons. In this respect, the present study extends the recent result of Herculano-Houzel [26] about the constancy of metabolic energy per neuron in the brains of mammals, by showing that also cerebral blood flow and capillary length per neuron are essentially conserved across species. There are approximately 10 µm of capillar- ies and 1.45 · 10−8 mL/min of blood flow per cortical neuron (Figs. 4 and 5; Supp. Tables S2 and S3). This finding suggests that not only brain metabolism but also its hemodynamics and microvascularization are evolutionarily constrained by the number of neurons. This mutual coupling might be a result of optimization in the design of cerebral energy expenditure and blood circulation. It should be underlined that both CBF and CMR scale with brain volume with the 13 exponent about −1/6, which is significantly different from the exponent −1/4 relating whole body resting specific metabolism with body volume [1, 2, 3]. Instead, the cerebral exponent −1/6 is closer to an exponent −0.12 ± 0.02 characterizing maximal body specific metabolic rate and specific cardiac output in strenuous exercise [43, 44]. In this sense, the brain metabolism and its hemodynamics resemble more the metabolism and circulation of exercised muscles than other resting organs, which is in line with the empirical evidence that brain is an energy expensive organ [10, 17, 18]. This may also suggest that there exists a common plan for the design of microcirculatory system in different parts of the mammalian body that uses the same optimization principles [45]. The results of this study show that as brain increases in size its capillary network becomes less dense, i.e. the densities of both capillary number and length decrease, respectively as Nc/V ∼ V −1/3−γ/24 and ρc ∼ V −1/6 (Table 3). Contrary to that, the capillary dimensions increase weakly with brain volume, their radius as Rc ∼ V 1/12 and their length segment as Lc ∼ V 1/6+γ/24, which are sufficient to make the fraction fc of capillary volume in the gray matter to be scale invariant (Table 3). The correction γ/24 appearing in the scaling exponents for Nc/V and Lc reflects the fact that blood viscosity depends on capillary radius (Fahraeus-Lindqvist effect [36]). This correction is however small for sufficiently large brains, generally for brains larger or equal to that of rat, for which typical values of γ/24 are in the range from −0.032 to −0.0004 (Table 2). On the contrary, for brains of mouse size or smaller, this correction is substantial, about −0.15, which implies that for very small brains Lc is essentially constant. Despite the changes in the geometry of microvessels, the velocity of capillary blood 14 uc is almost scale invariant for not too small brains (exponent γ/24 ≈ 0; Table 3). This prediction agrees with direct measurements of velocity in the brains of mouse, rat, and cat, which does not seem to change much, i.e. it is in the range 1.5 − 2.2 mm/sec [40, 46]. Consequently the transit time τc through a capillary increases with brain size as τc ∼ V 1/6, i.e. the scaling exponent is again 1/6. Another variable that seems to be independent of brain scale is partial oxygen pressure in cerebral capillaries (Table 3), which is consistent with the empirical findings in Fig. 3D on the invariance of oxygen pressure in arteries, as the two circulatory systems are mutually interconnected. 2. Capillary scaling in cerebral and non-cerebral tissue. The above scaling results for the brain can be compared with available analogous scaling rules for pulmonary, cardiovascular, and muscle systems. For these systems, it was proposed (no direct measurements) that partial oxygen pressure in capillaries should decline weakly with whole body volume (or organ volume as lung and heart volumes, Vlung, Vheart, scale isometrically with body volume [2]) with an exponent around −1/12, to account for the whole body specific metabolic exponent −1/4 [47, 48]. In the resting pulmonary system, the capillary radius as well as the density of capillary length scale the same way as they do in the brain, i.e., with the exponents 1/12 and −1/6, respectively, against system's volume [49]. Also, the capillary blood velocity in cerebral and non- cerebral tissues scale similarly, at least for not too small volumes, i.e. both are scale invariant [2, 3] (Table 3). However, the number of capillaries and capillary length seem to scale slightly different in the resting lungs, i.e. Nc ∼ V 5/8 lung and Lc ∼ V 5/24 lung [47], although the difference can be very mild. For the resting heart, it was predicted (again, 15 no direct measurements) that Nc ∼ V 3/4 heart, and blood transit time through a capillary τ ∼ V 1/4 heart [48], i.e. the exponents are multiples of a quarter power and are slightly larger than those for the brain (Table 3). Interestingly, for muscles and lungs in mammals exercising at their aerobic maxima, the blood transit time scales against body mass with an exponent close to 1/6 [50], which is the same as in the brain (Table 3). This again suggests that brain metabolism is similar to the metabolism of other maximally exercised organs. Overall, the small differences in the capillary characteristics among cerebral and non-cerebral resting tissues might account for the observed differences in the allometries of brain metabolism and whole body resting metabolism. In particular, the prevailing exponent 1/6 found in this study for brain capillaries, instead of 1/4, seems to be a direct cause for the distinctive brain metabolic scaling. 3. Brain microvascular network vs. neural network. The interesting question from an evolutionary perspective is how the allometric scalings for brain capillary dimensions relate to the allometry of neural characteristics. The neural density ρn (number of cortical neurons Nn per cortical gray matter volume V ) scales with cortical volume with a similar exponent as does the density of capillary length ρc (Fig. 4A). Thus, as a coarse-grained global description we have approximately ρn ∼ ρc (Fig. 4B,C), or Nn ∼ NcLc. The latter relation means that the total number of neurons is roughly proportional to the total length of capillaries, or equivalently, that capillary length per cortical neuron is conserved across different mammals. This cross-species conclusion is also in agreement with the experimental data for a single species. In particular, for mouse cerebral cortex it was found that densities of neural 16 number and microvessel length are correlated globally across cortical areas (but not locally within a single column) [51]. Moreover, since axons and dendrites occupy a constant fraction of cortical gray matter volume (roughly 1/3 each; [52, 53]), we have Nnld2 ∼ V , where l and d are respectively axon (or dendrite) length per neuron and diameter. Furthermore, because the average axon diameter d (unmyelinated) in the cortical gray matter is approximately invariant against the change of brain scale [52, 54], we obtain the following chain of proportionalities: l ∼ ρ−1 c , where n ∼ ρ−1 c ∼ V 1/6 ∼ Lα the exponent α = 1/(1 + γ/4). For medium and large brains, α ≈ 1, implying a nearly proportional dependence of axonal and dendritic lengths on capillary segment length. For very small brains (roughly below the volume of rat brain), α can be substantially greater than 1, suggesting a non-linear dependence between capillary and neural sizes. Given that the main exchange of oxygen between blood and brain takes place in the capillaries, these results suggest that metabolic needs of larger brains with greater but numerically sparser neurons must be matched by appropriately longer yet sparser capillaries. This finding reflects a rough, global relationship, which might or might not be related to the fact that during development neural and microvessel wirings share mutual mechanisms [20, 55]. At the cortical microscale, however, things could be more complicated, and a neuro-vascular correlation might be weaker, as both systems are highly plastic even in the adult brain (e.g. [56]). Regardless of its nature and precise dependence, the neuro-vascular coupling might be important for optimization of neural wiring [53, 57, 58]. In fact, neural connectivity in the cerebral cortex is very low, and it decreases with brain size [58, 59], similar to the density of capillary length (Fig. 3B, 17 Table 3). To make the neural connectivity denser, it would require longer axons and consequently longer capillaries. That may in turn increase excessively brain volume and its energy consumption, i.e. the costs of brain maintenance. As a result, the metabolic cost of having more neural connections and synapses for storing memories might outweigh its functional benefit. The brain metabolism is obviously strictly related to neural activities. In general, higher neural firing rates imply more cerebral energy consumed [18, 19]. It was esti- mated, based on a theoretical formula relating CMR with firing rate, that the latter should decline with brain size with an exponent around −0.15 [19]. This implies that neurons in larger brain are on average less active than neurons in smaller brains. Such sparse neural representations may be advantageous in terms of saving the metabolic en- ergy [18, 60, 61]. At the same time, what may be related, neural activity is distributed in such a way that both the average energy per neuron and the average blood flow per neuron are approximately invariant with respect to brain size (Fig. 5; Table 3, [26]). Additionally, average firing rate should be inversely proportional to the average blood transit time τc through a capillary, because both of them scale reversely with brain size (Table 3). Thus, it appears that global timing in neural activities should be correlated with the timing of cerebral blood flow. These general considerations suggest that apart from structural neuro-vascular coupling there is probably also a significant dynamic coupling. This conclusion is qualitatively compatible with experimental observations in which enhanced neural activity is invariably accompanied by increase in local blood flow [62]. 18 4. Relationship to brain functional imaging. The interdependencies between brain metabolism, blood flow, and capillary param- eters can have practical meaning. Currently existing techniques for non-invasive visu- alization of brain function, such as PET or fMRI, are associated with measurements of blood flow CBF and oxygen consumption CMRO2. It turns out that during stimulation of a specific brain region, CBF increases often, but not always, far more than CMRO2 [63]. However, both of them increase only by a small fraction in relation to the back- ground activity, even for massive stimulation [62, 63]. This phenomenon was initially interpreted as an uncoupling between blood perfusion and oxidative metabolism [64]. Later, it was shown that this asymmetry between CBF and CMRO2 can be explained in terms of mechanistic limitations on oxygen delivery to brain tissue through blood flow [65]. We can provide a related, but simpler explanation of these observations that involves physical limitations on the relative changes in capillary oxygen pressure and radius. During brain stimulation, both CBF and CMRO2 change by δCBF and δCMRO2, which are according to Eqs. (1) and (8) related to modifications in capillary radius (from Rc to Rc + δRc), and changes in partial oxygen pressure (pO2 7→ pO2 + δpO2). The density of perfused capillary length ρc remains constant for normal neurophysiological conditions. Accordingly, a small fraction of blood flow change is δCBF CBF ≈ (4 + γ) δRc Rc (9) 19 and similarly, a small fractional change in the oxygen metabolic rate is: δCMRO2 CMRO2 ≈ δpO2 pO2 . (10) In general, oxygen pressure increases with increasing capillary radius, in response to increase in blood flow CBF. This relationship can have a complicated character. We simply assume that pO2 ∼ Ra c , where the unknown exponent a (a > 0) contains all the non-linear effects, however complicated they are. Thus, a small fractional change in oxygen pressure can be written as δpO2/pO2 ≈ aδRc/Rc. As a result, we obtain δCMRO2 CMRO2 ≈ a (4 + γ) δCBF CBF . (11) If partial oxygen pressure pO2 depends on capillary radius linearly or sublinearly, i.e., if a ≤ 1, then the fractional increase in oxygen metabolism is significantly smaller than a corresponding increase in cerebral blood flow. This case corresponds to the experimental reports showing that this ratio is ≪ 1, for example, in the visual cortex (∼ 0.1) [66] and in the sensory cortex (∼ 0.2 − 0.4) [64, 67]. If, in turn, pO2 depends on Rc superlinearly, i.e. if a > 1, then the coefficient a/(4 + γ) in Eq. (10) can be of the order of unity. Such cases have been also reported experimentally during cognitive activities [42] or anesthesia [68, 69]. 20 Materials and Methods The ethics statement does not apply to this study. CBF data were collected from different sources: for mouse [70], rat [71], rabbit [72], cynomolgus monkey [73], rhesus monkey [74], pig [75], and human [76]. Cerebral capillary characteristics were obtained from several sources: for mouse [14, 51], rat [15, 77, 32], cat [40, 78], dog [79], rhesus monkey [80], and human [33, 34]. Data for calculating neuron densities were taken from [24, 25, 41, 52, 81, 82]. Cortical volume data (for 2 hemispheres) are taken from [52, 41, 81]. Their values are: mouse 0.12 cm3, rat 0.42 cm3, rabbit 4.0 cm3, cat 14.0 cm3, cynomolgus monkey 21.0 cm3, dog 35.0 cm3, rhesus monkey 42.9 cm3, pig 45.0 cm3, human 571.8 cm3. All the numerical data are provided in the Supporting Information (Tables S1, S2, and S3). 21 Supporting Information Appendix S1 Table S1 Regional cerebral blood flow CBF in mammals. Table S2 Cerebral capillary and neural characteristics in mammals. Table S3 Arterial partial oxygen pressure and average cortical CBF per neuron. 22 References [1] Kleiber M (1947) Body size and metabolic rate. Physiol. Rev. 27: 511- 541. [2] Schmidt-Nielsen K (1984) Scaling: why is animal size so important? Cambridge: Cambridge Univ. Press. [3] Calder WA (1984) Size, function, and life history. Cambridge, MA: Har- vard Univ. Press. [4] Dodds PS, Rothman DH, Weitz JS (2001) Re-examination of the '3/4- law' of metabolism. J. Theor. Biol. 209: 9-27. [5] Enquist BJ, West GB, Brown JH (2000) Quarter-power allometric scal- ing in vascular plants: functional basis and ecological consequences. In Scaling in biology, ed. Brown JH, West GB, pp. 113-128. Oxford: Oxford Univ. Press. [6] West GB, Brown JH, Enquist BJ (1997) A general model for the origin of allometric scaling laws in biology. Science 276: 122-126. [7] Banavar JR, Damuth J, Maritan A, Rinaldo A (2002) Supply-demand balance and metabolic scaling. Proc. Natl. Acad. Sci. USA 99: 10506- 10509. 23 [8] Darveau C, Suarez R, Andrews R, Hochachka P (2002) Allometric cas- cade as a unifying principle of body mass effects on metabolism. Nature 417: 166-170. [9] Savage VM, Deeds EJ, Fontana W (2008) Sizing up allometric scaling theory. PLoS Comput. Biol. 4(9): e1000171. [10] Karbowski J (2007) Global and regional brain metabolic scaling and its functional consequences. BMC Biology 5: 18. [11] Krogh A (1929) The anatomy and physiology of capillaries, 2nd ed. New Haven, CT: Yale Univ. Press. [12] Klein B, Kuschinsky W, Schrock H, Vetterlein F (1986) Interdependency of local capillary density, blood flow, and metabolism in rat brain. Am. J. Physiol. 251: H1333-H1340. [13] Borowsky IW, Collins RC (1989) Metabolic anatomy of brain: a com- parison of regional capillary density, glucose metabolism, and enzyme activities. J. Comp. Neurol. 288: 401-413. [14] Boero JA, Ascher J, Arregui A, Rovainen C, Woolsey TA (1999) In- creased brain capillaries in chronic hypoxia. J. Appl. Physiol. 86: 1211- 1219. [15] Hauck EF, Apostel S, Hoffmann JF, Heimann A, Kempski O (2004) Cap- illary flow and diameter changes during reperfusion after global cerebral 24 ischemia studied by intravital video microscopy. J. Cereb. Blood Flow Metab. 24: 383-391. [16] Ito H, Kanno I, Ibaraki M, Hatazawa J, Miura S (2003) Changes in hu- man cerebral blood flow and cerebral blood volume during hypercapnia and hypocapnia measured by positron emission tomography. J. Cereb. Blood Flow Metab. 23: 665-670. [17] Aiello LC, Wheeler P (1995) The expensive-tissue hypothesis: The brain and the digestive system in human and primate evolution. Curr. Anthro- pology 36: 199-221. [18] Attwell D, Laughlin SB (2001) An energy budget for signaling in the gray matter of the brain. J. Cereb. Blood Flow Metabol. 21: 1133-1145. [19] Karbowski J (2009) Thermodynamic constraints on neural dimensions, firing rates, brain temperature and size. J. Comput. Neurosci. 27: 415- 436. [20] Carmeliet P, Tessier-Lavigne M (2005) Common mechanisms of nerve and blood vessel wiring. Nature 436: 193-200. [21] Attwell D, Gibb A (2005) Neuroenergetics and the kinetic design of excitatory synapses. Nat. Rev. Neurosci. 6: 841-849. [22] Heeger DJ, Ress D (2002) What does fMRI tell us about neuronal ac- tivity? Nat. Rev. Neurosci. 3: 142-151. 25 [23] Logothetis NK, Wandell BA (2004) Interpreting the BOLD signal. Annu. Rev. Physiol. 66: 735-769. [24] Herculano-Houzel S, Mota B, Lent R (2006) Cellular scaling rules for rodent brains. Proc. Natl. Acad. Sci. USA 103: 12138-12143. [25] Herculano-Houzel S, Collins CE, Wong P, Kaas JH (2007) Cellular scal- ing rules for primate brains. Proc. Natl. Acad. Sci. USA 104: 3562-3567. [26] Herculano-Houzel S (2011) Scaling of brain metabolism with a fixed energy budget per neuron: Implications for neuronal activity, plasticity, and evolution. PLoS ONE 6(3): e17514. [27] Woodbury RA, Hamilton WF (1937) Blood pressure studies in small animals. Am. J. Physiol. 119: 663-674. [28] Gregg DE, Eckstein RW, Fineberg MH (1937) Pressure pulses and blood pressure values in unanesthetized dogs. Am. J. Physiol. 118: 399-410. [29] Li JK-J (2000) Scaling and invariants in Cardiovascular biology. In Scal- ing in biology, ed. Brown JH, West GB, pp. 113-128. Oxford: Oxford Univ. Press. [30] Lenigert-Follert E, Lubbers DW (1976) Behavior of microflow and local PO2 of the brain cortex during and after electrical stimulation. Pflugers Arch. 366: 39-44. 26 [31] Hakeem AY, Hof PR, Sherwood CC, Switzer RC, Rasmussen LEF, All- man JM (2005) Brain of the african elephant (Loxodonta africana): Neuroanatomy from magnetic resonance images. Anat. Rec. A 287 A: 1117-1127. [32] Michaloudi H, Batzios C, Grivas I, Chiotelli M, Papadopoulos GC (2006) Developmental changes in the vascular network of the rat visual areas 17, 18, and 18a. Brain Res. 1103: 1-12. [33] Meier-Ruge W, Hunziker O, Schulz U, Tobler HJ, Schweizer A (1980) Stereological changes in the capillary network and nerve cells of the aging human brain. Mechanisms of Ageing and Development 14: 233-243. [34] Lauwers F, Cassot F, Lauwers-Cances V, Puwanarajah P, Duvernoy H (2008) Morphometry of the human cerebral cortex microcirculation: General characteristics and space-related profiles. NeuroImage 39: 936- 948. [35] Sugihara-Seki M, Fu BM (2005) Blood flow and permeability in mi- crovessels. Fluid Dynamics Research 37: 82-132. [36] Fahraeus R, Lindqvist T (1931) The viscosity of the blood in narrow capillary tubes. Am. J. Physiol. 96: 562-568. [37] Pries AR, Secomb TW, Gaehtgens P (1996) Biophysical aspects of blood flow in the microvasculature. Cardiovasc. Res. 32: 654-667. 27 [38] Amin TM, Sirs JA (1985) The blood rheology of man and various animal species. Q. J. Exp. Physiol. 70: 37-49. [39] Schlageter KE, Molnar P, Lapin GD, Groothuis DR (1999) Microvessel organization and structure in experimental brain tumors: microvessel populations with distinctive structural and functional properties. Mi- crovascular Res. 58: 312-328. [40] Pawlik G, Rackl A, Bing RS (1981) Quantitative capillary topography and blood flow in the cerebral cortex of cat: an in vivo microscopic study. Brain Res. 208: 35-58. [41] Herculano-Houzel S, Mota B, Wong P, Kaas JH (2010) Connectivity- driven white matter scaling and folding in primate cerebral cortex. Proc. Natl. Acad. Sci. USA 107: 19008-19013. [42] Roland PE, Eriksson L, Stone-Elander S, Widen L (1987) Does mental activity change the oxidative metabolism of the brain? J. Neurosci. 7: 2373-2389. [43] Bishop CM (1999) The maximum oxygen consumption and aerobic scope of birds and mammals: getting to the heart of the matter. Proc. R. Soc. Lond. B 266: 2275-2281. [44] Weibel ER, Hoppeler H (2005) Exercise-induced maximal metabolic rate scales with muscle aerobic capacity. J. Exp. Biol. 208: 1635-1644. 28 [45] Weibel ER, Taylor CR, Hoppeler H (1991) The concept of symmorpho- sis: a testable hypothesis of structure-function relationship. Proc. Natl. Acad. Sci. USA 88: 10357-10361. [46] Unekawa M, Tomita M, Tomita Y, Toriumi H, Miyaki K, Suzuki N (2010) RBC velocities in single capillaries of mouse and rat brains are the same, despite 10-fold difference in body size. Brain Res. 1320: 69-73. [47] Dawson TH (2003) Scaling laws for capillary vessels of mammals at rest and in exercise. Proc. R. Soc. Lond. B 270: 755-763. [48] West GB, Brown JH, Enquist BJ (2000) The origin of universal scaling laws in biology. In Scaling in biology, ed. Brown JH, West GB, pp. 87- 112. Oxford: Oxford Univ. Press. [49] Dawson TH (2008) Modeling the vascular system and its capillary net- works. In Vascular Hemodynamics: Bioengineering and Clinical Per- spectives. Edited by Yim PJ. New York: Wiley. [50] Kayar SR, Hoppeler H, Jones JH, Longworth K, Armstrong RB, et al (1994) Capillary blood transit time in muscles in relation to body size and aerobic capacity. J. Exp. Biol. 194: 69-81. [51] Tsai PS, Kaufhold JP, Blinder P, Friedman B, Drew PJ, et al (2009) Correlations of neuronal and microvascular densities in murine cortex 29 revealed by direct counting and colocalization of nuclei and vessels. J. Neurosci. 29: 14553-14570. [52] Braitenberg V, Schuz A (1998) Cortex: Statistics and Geometry of Neu- ronal Connectivity. Berlin: Springer. [53] Chklovskii DB, Schikorski T, Stevens CF (2002) Wiring optimization in cortical circuits. Neuron 43: 341-347. [54] Olivares R, Montiel J, Aboitiz F (2001) Species differences and simi- larities in the fine structure of the mammalian corpus callosum. Brain Behav. Evol. 57: 98-105. [55] Stubbs D, DeProto J, Nie K, Englund C, Mahmud I, et al (2009) Neu- rovascular congruence during cerebral cortical development. Cereb. Cor- tex 19: 32-41. [56] Chklovskii DB, Mel BW, Svoboda K (2004) Cortical rewiring and infor- mation storage. Nature 431: 782-788. [57] Mitchison, G (1992) Neuronal branching patterns and the economy of cortical wiring. Proc. R. Soc. Lond. B 245: 151-158. [58] Karbowski J (2001) Optimal wiring principle and plateaus in the degree of separation for cortical neurons. Phys. Rev. Lett. 86: 3674-3677. 30 [59] Karbowski J (2003) How does connectivity between cortical areas de- pend on brain size? Implications for efficient computation. J. Comput. Neurosci. 15: 347-356. [60] Levy WB, Baxter RA (1996) Energy efficient neural codes. Neural Com- put. 8: 531-543. [61] Laughlin SB, de Ruyter van Steveninck RR, Anderson JC (1998) The metabolic cost of neural information. Nature Neurosci. 1: 36-41. [62] Moore CI, Cao R (2008) The hemo-neural hypothesis: on the role of blood flow in information processing. J. Neurophysiol. 99: 2035-2047. [63] Raichle ME, Mintun MA (2006) Brain work and brain imaging. Annu. Rev. Neurosci. 29: 449-476. [64] Fox PT, Raichle ME (1986) Focal physiological uncoupling of cerebral blood flow and oxidative metabolism during somatosensory stimulation in human subjects. Proc. Natl. Acad. Sci. USA 83: 1140-1144. [65] Buxton RB, Frank LR (1997) A model of the coupling between cerebral blood flow and oxygen metabolism during neural stimulation. J. Cereb. Blood Flow Metab. 17: 64-72. [66] Fox PT, Raichle ME, Mintun MA, Dence C (1988) Nonoxidative glucose consumption during focal physiologic neural activity. Science 241: 462- 464. 31 [67] Seitz RJ, Roland PE (1992) Vibratory stimulation increases and de- creases the regional cerebral blood flow and oxidative metabolism: a positron emission tomography (PET) study. Acta Neurol. Scand. 86: 60-67. [68] Nilsson B, Siesjo BK (1975) The effect of phenobarbitone anaesthesia on blood flow and oxygen consumption in the rat brain. Acta Anaesthesiol. Scand. Suppl. 57: 18-24. [69] Smith AL, Wollman H (1972) Cerebral blood flow and metabolism. Anesthesiology 36: 378-400. [70] Frietsch T, Maurer MH, Vogel J, Gassmann M, Kuschinsky W, et al (2007) Reduced cerebral blood flow but elevated cerebral glucose metabolic rate in erythropoietin overexpressing transgenic mice with excessive erythrocytosis. J. Cereb. Blood Flow Metab. 27: 469-476. [71] Frietsch T, Krafft P, Piepgras A, Lenz C, Kuschinsky W, et al (2000) Relationship between local cerebral blood flow and metabolism during mild and moderate hypothermia in rats. Anesthesiology 92: 754-763. [72] Tuor UI (1991) Local cerebral blood flow in the newborn rabbit: an autoradiographic study of changes during development. Pediatric Res. 29: 517-523. 32 [73] Orlandi C, Crane PD, Platts SH, Walovitch RC (1990) Regional cere- bral blood flow and distribution of [99mTc]Ethyl Cysteinate dimer in nonhuman primates. Stroke 21: 1059-1063. [74] Noda A, Ohba H, Kakiuchi T, Futatsubashi M, Tsukada H, et al (2002) Age-related changes in cerebral blood flow and glucose metabolism in conscious rhesus monkeys. Brain Res. 936: 76-81. [75] Delp MD, Armstrong RB, Godfrey DA, Laughlin MH, Ross CD, et al (2001) Exercise increases blood flow to locomotor, vestibular, cardiores- piratory and visual regions of the brain in miniature swine. J. Physiol. 533.3 849-859. [76] Bentourkia M, Bol A, Ivanoiu A, Labar D, Sibomana M, et al (2000) Comparison of regional cerebral blood flow and glucose metabolism in the normal brain: effect of aging. J. Neurol. Sci. 181: 19-28. [77] Bar TH (1980) The vascular system of the cerebral cortex. Adv. Anat. Embryol. Cell Biol. 59: 71-84. [78] Tieman SB, Moller S, Tieman DG, White JT (2004) The blood supply of the cat's visual cortex and its postnatal development. Brain Res. 998: 100-112. 33 [79] Luciano MG, Skarupa DJ, Booth AM, Wood AS, Brant CL, et al (2001) Cerebrovascular adaptation in chronic hydrocephalus. J. Cereb. Blood Flow Metab. 21: 285-294. [80] Weber B, Keller AL, Reichold J, Logothetis NK (2008) The microvascu- lar system of the striate and extrastriate visual cortex of the macaque. Cereb. Cortex 18: 2318-2330. [81] Mayhew TM, Mwamengele GLM, Dantzer V (1996) Stereological and allometric studies on mammalian cerebral cortex with implications for medical brain imaging. J. Anat. 189: 177-184. [82] Haug H (1987) Brain sizes, surfaces, and neuronal sizes of the cortex cerebri: A stereological investigation of Man and his variability and a comparison with some mammals (primates, whales, marsupials, insecti- vores, and one elephant). Am. J. Anatomy 180: 126-142. [83] Hatazawa J et al (1995) Regional cerebral blood flow, blood volume, oxygen extraction fraction, and oxygen utilization rate in normal volun- teers measured by the autoradiographic technique and the single breath inhalation method. Ann. Nucl. Med. 9: 15-21. [84] Heistad DD et al (1976) Effect of stimulation of carotid chemoreceptors on total and regional cerebral blood flow. Cir. Res. 38: 20-25. 34 Figure Captions Fig. 1 Scaling of cerebral blood flow CBF in the cortical gray matter. (A) Visual cortex: y = −0.127x + 1.98 (R2 = 0.93, p < 0.001). 95% confidence interval for the slope CI=(- 0.168,-0.086). (B) Parietal cortex: y = −0.150x + 2.00 (R2 = 0.89, p = 0.005), slope CI=(-0.222,-0.078). (C) Frontal cortex: y = −0.170x + 2.00 (R2 = 0.89, p = 0.002), slope CI=(-0.239,-0.100). (D) Temporal cortex: y = −0.191x + 2.09 (R2 = 0.89, p = 0.005), slope CI=(-0.286,-0.096). Fig. 2 Scaling of cerebral blood flow CBF in the subcortical gray matter. (A) Hippocampus: y = −0.135x + 1.89 (R2 = 0.90, p = 0.049), slope CI=(-0.271,0.000). (B) Thalamus: y = −0.167x + 2.02 (R2 = 0.83, p = 0.011), slope CI=(-0.272,-0.062). (C) Cerebellum: y = −0.177x + 2.03 (R2 = 0.88, p = 0.002), slope CI=(-0.252,-0.102). Fig. 3 Scaling of brain capillary characteristics against brain size. (A) Capillary diameter scales against cortical gray matter volume with the exponent 0.075 (y = 0.075x + 0.60, R2 = 0.87, p = 0.007), exponent CI=(0.034,0.117). (B) Volume density of capillary length ρc scales with the exponent −0.16 (y = −0.162x + 2.86, R2 = 0.79, p = 0.044), exponent CI=(-0.316,-0.008). (C) Fraction of capillary volume fc in gray matter is essentially independent of brain size (y = 0.029x−1.83, R2 = 0.07, p = 0.662), the same as (D) the arterial partial oxygen pressure (y = 0.012x + 1.96, R2 = 0.09, p = 0.472). 35 Fig. 4 Neuron density versus capillary length density in the cerebral cortex. (A) Across our sample of mammals, the cortical neuron number density ρn scales against cortical vol- ume with the exponent −0.13 (y = −0.128x + 1.86, R2 = 0.87, p = 0.022), exponent CI=(-0.221,-0.036). (B) The ratio of the density of capillary length ρc to the density of neurons ρn in the cortex does not correlate with brain size (y = −0.034x + 0.99, R2 = 0.09, p = 0.617), exponent CI=(-0.228,0.161). (C) The log-log dependence of the capillary length density ρc on neuron density ρn gives the exponent of 1.05 (y = 1.051x + 0.88, R2 = 0.63, p = 0.109). Fig. 5 Invariance of cerebral blood flow per cortical neuron across mammals. The ratio of CBF to neuron density ρn in the cerebral cortex does not correlate significantly with brain size (log-log plot yields y = −0.033x + 0.183, R2 = 0.39, p = 0.261). The value of CBF for each species is the arithmetic mean of regional CBF across cerebral cortex. 36 Table 1: Parameters affecting the effective blood viscosity. Rc [µm] w [µm] w/Rc 0.05 1.5 0.15 2.0 0.24 2.5 3.0 0.30 0.07 0.30 0.60 0.90 1 − (1 − µ)(1 − w/Rc)4 0.27 0.54 0.71 0.79 0.85(ln(Rc/1.2))2/3 0.31 0.54 0.69 0.80 Data for w and Rc were collected from [35]. The value of µ was taken as 1/8. The last column represents values of the fitting function to the function in the fourth column. Table 2: Exponent γ as a function of capillary diameter. dog monkey human Species 4.5 2Rc [µm] γ -0.49 mouse 3.1 -3.51 6.4 -0.01 rat 4.1 -0.77 cat 5.1 -0.25 5.6 -0.13 Data for Rc were taken from Suppl. Inform. Table S2 (references therein). Table S1: Regional cerebral blood flow CBF in mammals. Species Mouse (a) Rat (b) Rabbit (c) Cynomolgus monkey (d) Rhesus monkey (e) Pig (f ) Human (g) visual frontal cortex cortex 165.0 124.0 124.0 116.0 69.6 67.2 temporal parietal hippocam. cortex 157.0 195.0 -- cortex 166.0 113.0 63.6 112.8 77.8 -- 55.4 59.2 66.5 43.0 48.7 45.0 63.3 41.2 70.6 53.2 45.5 44.6 56.8 -- 62.8 43.4 -- 53.9 41.0 -- thalamus cerebell. 195.5 116.0 60.6 56.7 -- 48.5 47.5 195.0 123.0 62.4 57.2 47.8 64.0 41.5 CBF data are given in mL/(100g*min). References: (a) [70]; (b) [71]; (c) [72]; (d) [73]; (e) [74]; (f ) [75]; (g) [76]. 37 Table 3: Summary of scalings for brain capillaries and hemodynamics against cortical gray matter volume V . Parameter Capillary radius, Rc Capillary length density, ρc Capillary volume fraction, fc Total capillary number, Nc Capillary segment length, Lc Capillary blood velocity, uc Capillary transit time, τc Capillary oxygen pressure, pO2 Capillary length per neuron, ρc/ρn Cerebral blood flow, CBF Blood flow per neuron, CBF/ρn Oxygen consumption rate, CMRO2 CMRO2 ∼ V −1/6 Oxygen use per neuron, CMRO2/ρn CMRO2/ρn ∼ V 0 Scaling rule Rc ∼ V 1/12 ρc ∼ V −1/6 fc ∼ V 0 Nc ∼ V 2/3−γ/24 Lc ∼ V 1/6+γ/24 uc ∼ V γ/24 τc ∼ V 1/6 pO2 ∼ V 0 ρc/ρn ∼ V 0 CBF ∼ V −1/6 CBF/ρn ∼ V 0 38 Table S2: Cerebral capillary and neural characteristics in mammals. ρc/ρn Species [µm] 7.59 11.29 15.60 0.015 (b) 0.014 (i) 0.021 (d) fc 2Rc [µm] 3.1 (a,b) 4.1 (c) 5.1 (d) 4.5 (e) -- Mouse Rat Cat Dog Rhesus monkey 5.6 (f ) Human 6.4 (g,h) 0.009 (f ) 0.023 (g) ρc [mm/mm3] 880.0 (a,b) 806.0 (i) 780.0 (j) -- 329.9 (f ) 219.0 (g) ρn [106/cm3] 116.0 (k) 71.4 (l) 50.0 (m) 12.6 (m) -- 37.4 (n) 38.5 (n) 8.82 5.69 The ratio ρc/ρn is the average capillary length per neuron. References: (a) [14]; (b) [51]; (c) [32]; (d) [40]; (e) [79]; (f ) [80]; (g) [33]; (h) [34]; (i) [77]; (j) [78]; (k) [52]; (l) [24]; (m) [82]; (n) [41]. 39 Table S3: Arterial partial oxygen pressure and average cortical CBF. arterial pO2 [mm Hg] CBFcortex CBFcortex/ρn 93.0 (a) 93.0 (b) 75.0 (c) 157.4 (a) 131.8 (b) 70.2 (c) Species Mouse Rat Rabbit Cynomolgus 102.4 (d) monkey 107.0 (e) Dog Rhesus monkey 92.4 (f ) 104.0 (g) Pig 94.0 (h,i) Human 1.36 1.84 -- 57.9 (d) -- 52.5 (f ) 59.5 (g) 42.6 (j) 1.51 -- 1.40 -- 1.11 The ratio CBFcortex/ρn is the average cortical blood flow per neuron (10−8 mL/min). References: (a) [70]; (b) [71]; (c) [72]; (d) [73]; (e) [84]; (f ) [74]; (g) [75]; (h) [16]; (i) [83]; (j) [76]. 40 Appendix S1 In this Appendix we want to show a formal step connecting Eq. (6) and Eq. (8) in the main text. We want to find the exponent γ, which satisfies the following equation: (ln(Rc/Ro))2/3 = (Rc/Ro)γ. Taking the natural logarithm of both sides leads to (2/3) ln(ln(Rc/Ro)) = γ ln(Rc/Ro), which yields Eq. (7) in the main text. 41 y = − 0.127*x + 1.97 i ] ) n m * g 0 0 1 ( / L m [ ) x e t r o c l a t e i r a p F B C ( 0 1 g o L y = − 0.15*x + 2 B 2.4 2.3 2.2 2.1 2 1.9 1.8 1.7 1.6 1.5 A 2.4 2.3 2.2 2.1 2 1.9 1.8 1.7 1.6 1.5 1.4 −2 −1 0 i ] ) n m * g 0 0 1 ( / L m [ ) x e t r o c l i a u s v F B C ( g o L 0 1 C 2.4 2.3 2.2 2.1 2 1.9 1.8 1.7 1.6 1.5 1.4 −2 −1 0 i ] ) n m * g 0 0 1 ( / L m [ ) x e t r o c l a t n o r f F B C ( g o L 0 1 1 2 Log 10 (cortical volume) [cm3] 3 4 1.4 −2 −1 0 1 2 3 4 Log 10 (cortical volume) [cm3] y = − 0.17*x + 2 i ] ) n m * g 0 0 1 ( / L m [ ) x e t r o c l a r o p m e t F B C ( 0 1 g o L y = − 0.19*x + 2.1 D 2.4 2.3 2.2 2.1 2 1.9 1.8 1.7 1.6 1.5 1 2 Log 10 (cortical volume) [cm3] 3 4 1.4 −2 −1 0 1 2 3 4 Log 10 (cortical volume) [cm3] Figure 1 A 2.3 2.2 2.1 2 1.9 1.8 1.7 1.6 1.5 1.4 −2 −1 i ] ) n m * g 0 0 1 ( / L m [ ) s u p m a c o p p h F B C i ( g o L 0 1 y = − 0.14*x + 1.9 i ] ) n m * g 0 0 1 ( / L m [ ) s u m a a h t l F B C ( 0 1 g o L y = − 0.17*x + 2 B 2.4 2.2 2 1.8 1.6 0 1 Log 10 (cortical volume) [cm3] C 2 3 1.4 −2 −1 0 1 2 3 4 Log 10 (cortical volume) [cm3] y = − 0.18*x + 2 i ] ) n m * g 0 0 1 ( / L m [ ) m u l l e b e r e c F B C ( g o L 0 1 2.4 2.2 2 1.8 1.6 1.4 −2 −1 0 Figure 2 1 2 3 4 Log 10 (cortical volume) [cm3] ] m µ [ ) r e t e m a d y r a i l l i p a c ( 0 1 g o L A 1 0.9 0.8 0.7 0.6 0.5 0.4 −2 C B 3.2 3 2.8 2.6 2.4 2.2 −2 D −1 y = 0.075*x + 0.6 ] 3 m m m m / [ ) ρ ( c 0 1 g o L y = − 0.162*x + 2.86 −1 0 1 2 Log 10 (cortical volume) [cm3] 1 0 (cortical volume) [cm3] 2 Log 10 3 4 3 4 y = 0.029*x − 1.8 −1 −1.5 −2 −2.5 l ) e m u o v y r a l l i p a c f o n o i t c a r f ( g o L 0 1 −3 −2 −1 1 0 (cortical volume) [cm3] 2 Log 10 3 4 3 4 ] g H m m [ ) e r u s s e r p 2 O l a i r e t r a ( 0 1 g o L y = 0.012*x + 2 3 2.8 2.6 2.4 2.2 2 1.8 1.6 1.4 1.2 1 −2 −1 0 1 2 Log 10 (cortical volume) [cm3] Figure 3 A 2.2 2 1.8 1.6 1.4 1.2 −2 −1 0 ] 3 m c / 6 0 1 [ ) y t i s n e d n o r u e n ( 0 1 g o L y = − 0.13*x + 1.9 ] m µ [ ) ρ / ρ ( n c 0 1 g o L B 3 2.5 2 1.5 1 0.5 0 −0.5 −1 −2 y = − 0.034*x + 0.99 −1 0 1 2 3 4 Log 10 (cortical volume) [cm3] 1 2 3 4 Log 10 (cortical volume) [cm3] y = 1.05*x + 0.875 C 3.2 3 2.8 2.6 2.4 2.2 ] 3 m m m m / [ ) c ρ ( 0 1 g o L 2 1.4 1.5 1.6 1.7 1.9 1.8 2 (ρ ) [106/cm3] n 10 Log 2.1 2.2 Figure 4 y = − 0.033*x + 0.18 1.5 1 0.5 0 −0.5 ) n ρ / F B C ( 0 1 g o L −1 −2 −1 0 Log 10 (cortical volume) [cm3] 1 2 3 4 Figure 5
1312.1104
1
1312
2013-12-04T10:57:48
New insights in gill/buccal rhythm spiking activity and CO2 sensitivity in pre- and post-metamorphic tadpoles (Pelophylax ridibundus)
[ "q-bio.NC", "physics.bio-ph" ]
Central CO2chemosensitivity is crucial for all air-breathing vertebrates and raises the question of itsrole in ventilatory rhythmogenesis. In this study, neurograms of ventilatory motor outputs recorded infacial nerve of premetamorphic and postmetamorphic tadpole isolated brainstems, under normo- andhypercapnia, are investigated using Continuous Wavelet Transform spectral analysis for buccal activityand computation of number and amplitude of spikes during buccal and lung activities. Buccal burstsexhibit fast oscillations (20-30 Hz) that are prominent in premetamorphic tadpoles: they result from thepresence in periodic time windows of high amplitude spikes. Hypercapnia systematically decreases thefrequency of buccal rhythm in both pre- and postmetamorphic tadpoles, by a lengthening of the interburstduration. In postmetamorphic tadpoles, hypercapnia reduces buccal burst amplitude and unmasks smallfast oscillations. Our results suggest a common effect of the hypercapnia on the buccal part of the CentralPattern Generator in all tadpoles and a possible effect at the level of the motoneuron recruitment inpostmetamorphic tadpoles.
q-bio.NC
q-bio
1 New insights in gill/buccal rhythm spiking activity and CO2 sensitivity in pre- and postmetamorphic tadpoles (Pelophylax ridibundus) Brigitte QUENETa, Christian STRAUS b, c, Marie-Noëlle FIAMMAb, Isabelle RIVALS a, Thomas SIMILOWSKI b, d, Ginette HORCHOLLE-BOSSAVIT a a. ESPCI-ParisTech, Equipe de Statistique Appliquée, F-75005 Paris, France b. UPMC Univ Paris 06, ER 10 UPMC, F-75013 Paris, France c. Assistance Publique - Hôpitaux de Paris, Groupe Hospitalier Pitié-Salpêtrière, Service Central d'Explorations Fonctionnelles Respiratoires, F-75013 Paris, France d. Assistance Publique - Hôpitaux de Paris, Groupe Hospitalier Pitié-Salpêtrière, Service de Pneumologie et Réanimation Médicale, F-75013 Paris, France Corresponding author at: ESPCI-ParisTech, Equipe de Statistique Appliquée, F-75005 Paris, France. Tel.: +33 140794461; Fax: +33 1 40794669 E-mail address [email protected] (B Quenet). ABSTRACT Central CO2 chemosensitivity is crucial for all air-breathing vertebrates and raises the question of its role in ventilatory rhythmogenesis. In this study, neurograms of ventilatory motor outputs recorded in facial nerve of premetamorphic and postmetamorphic tadpole isolated brainstems, under normo- and hypercapnia, are investigated using Continuous Wavelet Transform spectral analysis for buccal activity and computation of number and amplitude of spikes during buccal and lung activities. Buccal bursts exhibit fast oscillations (20-30Hz) that are prominent in premetamorphic tadpoles: they result from the presence in periodic time windows of high amplitude spikes. Hypercapnia systematically decreases the frequency of buccal rhythm in both pre- and postmetamorphic tadpoles, by a lengthening of the interburst duration. In postmetamorphic tadpoles, hypercapnia reduces buccal burst amplitude and unmasks small fast oscillations. Our results suggest a common effect of the hypercapnia on the buccal part of the Central Pattern Generator in all tadpoles and a possible effect at the level of the motoneuron recruitment in postmetamorphic tadpoles. KEYWORDS Amphibian buccal ventilation, central CO2 chemosensitivity, fast oscillations, CWT spectrum, spiking activity. 2 1. Introduction in amphibians, characterized by a the breathing control Previous studies of developmental transition from water breathing to air breathing, have addressed the issue of the emergence of central CO2 chemosensitivity. Many investigations use the fact that cranial nerve outputs of the isolated brainstem preparations from tadpoles and frogs match the corresponding respiratory muscle sequences (Gdovin et al., 1998; Kogo et al., 1994a, 1994b; Sakakibara, 1984a, 1984b). The nerve activity exhibits regular low amplitude bursts associated with gill ventilation and episodic high amplitude bursts corresponding to lung ventilation (Galante et al., 1996). Several studies have shown that hypercapnia increases systematically the occurrence of lung episodes in postmetamorphic tadpoles (Kinkead and Milsom, 1994; Torgerson et al., 1997; Taylor et al., 2003a, 2003b; Wilson et al., 2002; Straus et al., 2011). But regarding the buccal rhythm, experimental results describe heterogeneous effects of hypercapnia and metamorphosis on the gill/buccal frequency in pre- and postmetamorphic tadpoles, leading to contradictory conclusions. In Lithobates catesbeianus (Rana catesbaiana) of the same premetamorphic stages exposed to hypercapnia, Torgerson et al. (1997) reported an increase in the frequency and the amplitude of the gill rhythm while Taylor et al. (2003a) observed a tendency for frequency to decrease and no change in duty cycle. In postmetamorphic animals of the same stage exposed to hypercapnia, Torgerson et al. (1997) did not observe any response; Taylor et al. (2003b) observed a tendency for frequency to decrease. These conflicting conclusions could be explained by physiological differences between species or/and by methodological differences. Several studies reported effect of metamorphosis at normocapnia in Lithobates catesbeianus (Burggren and Doyle, 1986; Togerson et al., 1998; Gdovin et al., 1999; Taylor et al., 2003b; Hedrick et al., 2005; Fournier and Kinkead, 2006; Chen and Hedrick, 2008). These studies gave conflicting results about the gill/buccal activity: for instance, Taylor et al. (2003b) and Fournier and Kinkead (2006) reported no change of the buccal frequency, while Chen and Hedrick (2008) observed a decrease of about 50%. Developmental observations in tadpoles (Torgerson et al., 2001a, 2001b; Taylor et al., 2003b) have suggested that the neural substrates for central chemoreception exist in the vicinity of the structures responsible for both gill/buccal and lung rhythm generation already in premetamorphic tadpoles. Here, we study how CO2 sensitivity affects premetamorphic and postmetamorphic gill/buccal and lung rhythm in Pelophylax ridibundus. We analyze neurograms recorded from isolated brainstems in Pelophylax ridibundus whose chaotic nature has been evidenced by Straus et al. (2011). On the part of the filtered neurograms corresponding to buccal activity only, we perform time-frequency transform with complex Morlet wavelets (CWT) that is adapted to non-stationary signals (Akay, 2005; Mor and Lev- Tov, 2007; Vialatte et al., 2007; Romcy-Pereira et al., 2008). With adapted signal-processing tools, we evidenced, for the first time in amphibians, the presence of fast oscillations in the 20- 30 Hz frequency band in the buccal bursts, probably resulting from phase locked motoneuron activities, which are reminiscent of some of the fast oscillations described in several species of mammals (Funk and Parkis, 2002; Marchenko and Rogers, 2006, 2007). Regarding the effects of hypercapnia (acidosis at pH=7.4) on the gill/buccal activity, our results show that there is no change of the fast oscillations in the premetamorphic tadpole, but significant changes in postmetamorphic ones. In addition, hypercapnia induces significant lengthening of the buccal interburst period at both stages. 3 2. Materials and methods 2.1. Data collection Using the brainstem preparation of the tadpole, Pelophylax ridibundus (Rana esculenta - European Edible Frog) described in detail in Straus et al. (2011), the present study investigates a subset of the signals used in this paper. The experimental protocol fulfilled appropriate legal and ethical conditions (French Ministry for Agriculture and Animal Care Committee - Ile-de-France, Paris, Comité 3). Briefly, the developmental stage of the animals is determined according to Taylor and Köllros (1946). Tadpoles are chosen in order to assign them either to the premetamorphic (stages X-XIII) or to the postmetamorphic (stage XXIV- XXV) group of development. They are anesthetized and decerebrated. The brainstem is dissected and transferred to a recording chamber. This chamber is superfused at room temperature (~21°C) with mock cerebrospinal fluid (CSF) equilibrated with a gas mixture of O2 and CO2: (in mM) NaCl, 104; KCl, 4; MgCl2, 1.4; D-glucose, 10; NaHCO3, 25; CaCl2 2.4. The pH of the CSF is set to 7.8 or 7.4 by adjusting the fractional concentration of CO2. The electrical activity – namely the neurogram - of the root of the facial nerve (VII) is recorded using glass suction electrodes. The raw signal is then amplified (10 000 times) and filtered (100 Hz–1 kHz) using a high gain, differential AC amplifier (model 1700, A-M Systems Inc., Evrett, WA), digitized at 2000 Hz (PowerLab, ADInstruments) and stored as text files. Recordings last 20-35 min for each pH value. The first 10 minutes of each recording is discarded. Then, 5 minutes of signal with the best identifiable bursts of buccal and lung ventilation are selected for analysis. Indeed, the tadpole isolated brainstem preparation exhibits neural ventilatory activity but also periods of apnea and, sometimes, also bursts of undetermined kind, like, for example, the so-called ‘complex discharges’ (Straus et al., 2000). We retain the part of the neurograms where the preparation is not apneic at that time the type of the bursts is clear, either buccal or lung, with a minimum number of discharges of an undeterminable kind. 2.2. Neurogram processing The analysis is performed on neurograms from 5 premetamorphic (T, 1-5) and 5 postmetamorphic (F, 1-5) preparations, in normo- (pH=7.8) and hypercapnia (acidosis at pH=7.4). The spectral analysis based on complex Morlet wavelets (CWT) leads to information in the frequency domain taking temporal short-time variations into account. After segmentation of the signal into cycles (buccal and lung), we identified the lung bursts, and computed cross-correlation functions in order to insure the best temporal alignment of these segmented cycles of each type, buccal or lung, separately. Finally, we come back to the neurograms, which can be analyzed using the previous segmentation and temporal alignment of the cycles, in order to measure the position and amplitude of each spike in each cycle. 2.3. Neurogram filtering and spike detection Each neurogram is squared, zero-phase filtered (Gustafsson, 1996) using a moving average rectangular window of width W =100 ms. After taking the square root of the result, the lung bursts are eliminated (see next paragraph). On the resulting S100 signal, a detection of local minima is performed in order to segment this buccal signal into cycles, preserving both the buccal burst and the inactive period in each cycle (Fig. 1A, black line). The same filter with a window of width W=10 ms is applied to the buccal part of the neurogram, Sneur (Fig. 1A, gray line) resulting in a new signal S10 (Fig. 1A, red line). 4 A detection of maxima, i.e. spikes, is performed independently on the positive and the negative parts, in absolute value, on Sneur (Fig. 1B, red and blue dots), in order to store the time position and the amplitude of each of them. This detection allows the computation of a local spike density (per time unit). 2.4. Lung burst detection and analysis The detection of local minima performed on the squared and filtered neurogram (with W=100 ms) is completed by a detection of each local maximum between two successive minima defining each burst. In a given neurogram, a burst is labeled ‘lung burst’ according to two criteria: its amplitude is higher than twice the mean amplitude of all the recorded burst, and its spike density is higher than twice the mean spike density of all the recorded bursts. 2.5. Cycle time alignment by cross-correlation function The aim of this signal-processing step is to get a time alignment of the buccal bursts able to exhibit phase locking of neuronal activity inside the bursts of a given neurogram. Therefore, we define in S10 a “reference cycle” and position all the other cycles according to the maximal value of the cross-correlation function (sliding dot product) between each cycle and the reference (Lynn and Fuerst, 2000). At a first step, the amplitude vectors representing the S10 buccal cycles, after segmentation, are padded with zeros when necessary, in order to get an identical length for all these vectors. In a second step, a symmetric matrix of correlations is computed, whose generic term of indices i and j is the maximal value of the cross-correlation function between the amplitude vectors of cycle i and cycle j. The “reference cycle” is then the cycle whose mean correlation with all the other cycles is the highest: it is therefore considered as the most representative cycle. At a third step, the time lag, which corresponds to the maximum of the cross-correlation function between a cycle and the reference cycle, defines the time position of this cycle relatively to the reference (Fig. 2A): when the cycles are positioned according to their time lag and ranked according to their correlation with the reference, the bursts are aligned, and each cycle can be visualized in a map by a horizontal band where its amplitude is coded in gray scale (Fig. 2A). A same time alignment procedure is applied to the lung burst. 2.6. Amplitude profiles from the “spike matrices” The aim of this procedure is to get canonical profiles of the burst and interburst part representative of each neurogram buccal cycles. Using the relative time positions defined previously, we build two matrices of the cycle positive and negative spikes respectively (Fig 1B). For each matrix, the generic term of indices i and j is the amplitude of a spike, when present, at sampling time j of cycle i. In maps visualizing these matrices, each row corresponds to a cycle, each column corresponds to a time bin of the cycles, and a dot in a row and a column corresponds to a spike in that time bin for this cycle (Fig 2B). An amplitude profile is computed from these two matrices: the mean values, at each time bin, of the positive and negative amplitudes over the cycles, after zero-phase filtering with a window of 5 ms, constitute the two parts of the profile. In order to define a canonical cycle for each neurogram, characterized by its amplitude profile, with its positive and negative parts, whose duration corresponds to the median value of the cycle durations, we use the most representative part of the amplitude profiles, i.e. where the number of zeros related to the padding is the smallest: this criterion defines the beginning and the end of the canonical cycle (Fig. 2A and 2B). In such a canonical cycle, we want to define a central active zone, or duty cycle, corresponding to the buccal burst, surrounded by zones of minor activity. The beginning of the burst is the first time 5 when the positive amplitude profile is higher than its mean value; symmetrically, the end of the burst is the last time when the positive amplitude profile is higher that its mean value. As the positive and the negative amplitude profiles are very similar in absolute value, only the positive part is visualized when comparing the canonical cycles at the two values of the pH; moreover, in graphical representations of a canonical activity, the positive part of the canonical cycle is repeated twice, in order to get a complete view of the interburst part. Amplitude profiles of lung burst are computed using the same procedure. 2.7. Number and amplitude profiles according to spike amplitude The aim of this procedure is to evaluate the contributions of the spikes to the time structure of the bursts according to their amplitude. The alternation of high and weak spike activity characterizes the cyclic buccal part of the neurogram: we define the high activity window (buccal burst duration or ‘duty cycle’) as the time interval of the buccal canonical cycle where its profile amplitude is higher than the mean amplitude of this cycle (between the vertical green lines in Fig. 2B). Then, two histograms of spike amplitude are computed in the buccal cycles -aligned according to the previously described cross-correlation procedure (see § 2.5)- corresponding to the spikes inside the active window (Pin) and to the spikes outside (Pout) this window respectively. Populations Pin and Pout exhibit lognormal-like shaped distributions (not shown) of the spike amplitudes. The threshold amplitude value Ath where the number of spikes of Pin becomes higher that the number of spikes of Pout, defines the minimal amplitude of a spike to have a higher probability to belong to Pin than to Pout. The positive spike population is then divided into five subpopulations. The first one corresponds to the spikes whose amplitude is smaller than Ath, the second, third and fourth subpopulations contain spikes whose amplitude belongs to ]Ath, 3*Ath], ]3*Ath, 5*Ath] and ]5*Ath, 7*Ath] respectively, and the fifth subpopulation contains the highest remaining spikes, higher than 7*At (Fig. 1B). For each subpopulation k (k=1, 2, 3, 4 or 5), two new matrices are computed: a matrix of the number of spikes over ten consecutive time bins Mk and a matrix of the mean amplitude of these spikes Pk. In order to visualize the time distribution of the spiking activity for each subpopulation, Mk is graphically represented by a map of dots, a dot corresponding to the presence of one spike at least (Fig. 2B). The time successive mean values of the number of spikes over the cycles define a number profile for each subpopulation, idem for the time successive mean values of the amplitude, which define the amplitude profile for each subpopulation. 2.8. Continuous Wavelet Transform maps and spectra Given the non-stationarity of the buccal rhythm, a time-frequency analysis of S10 signal is performed using a Continuous Wavelet Transform (CWT) (Akay, 2005; Mor & Lev-Tov; 2007; Vialatte et al., 2007; Romcy-Pereira et al., 2008) of S10 subsampled at 200 Hz (after applying an anti-aliasing low pass filter eliminating all the frequencies above 100 Hz), looking for relevant frequencies between 1 and 100 Hz, with a complex Morlet wavelet basis function including five oscillations (Mallat, 2000). The coefficients of each time- frequency amplitude map are computed for 100 values logarithmically distributed between 1 and 100 Hz, and at a time resolution corresponding to the sampling of S10. This time- frequency map displays the temporal evolution of the S10 dominant frequencies in gray scale (Fig. 3). All the CWT maps got from our signals exhibit a relative stability of the dominant frequencies, which validates the representation of each S10 by a CWT spectrum computed as the mean value, at each frequency, of the corresponding coefficients of the amplitude map over time activity (Fig. 3). In order to compare these spectra, each of them is normalized. 6 2.9. Statistical analysis Statistical tests have been applied to the three following descriptors: the mean buccal cycle duration, the relative energy in the band 1-10 Hz and the relative energy in the band 20-30 Hz, measured on the CWT spectrum. In order to test both the effects of pH and stage on these descriptors, a two-way ANOVA is applied when the condition of homoscedasticity is fulfilled, which is tested with Bartlett’s test. In this case, we also have to take into account the fact that the same animal is tested for the two values of the pH, by including a random individual effect. When homoscedasticity is rejected, pairwise statistical tests are performed. The effect of the pH at a given stage is tested with a paired t-test, the effect of the stage at a given pH is tested with a standard t-test if the variance of the two groups are homogeneous (F-test), with a Wilcoxon rank sum test if not. Statistical significance was assumed if p < 0.05. Signal processing algorithms and statistical computations have been implemented with Matlab (The Mathworks, version 8.0.0.783 (R2012b)). 3. Results 3.1. CWT maps and spectra of buccal activity Frequency information from S10 signal is visualized in both CWT maps and spectra, as illustrated for T2 and F4 at both values of the pH (Fig. 3, lower part). Three seconds of the corresponding S10 signals are illustrated in the upper part of this figure. In time domain, the main effects induced by metamorphosis and by acidosis on the buccal rhythm are already visible, i.e. a shortening of the buccal period after metamorphosis, a lengthening of the period with acidosis in both premetamorphic and postmetamorphic tadpoles, and oscillations superimposed on the buccal bursts at both pH in premetamorphic tadpoles, much more discrete in postmetamorphic tadpoles. For T2, the maps reveal a dark gray strip, discontinuous, at low frequency, 1.5 Hz, corresponding to the fundamental buccal frequency at pH=7.8, whereas the equivalent strip is darker, continuous, at 1.2 Hz at pH=7.4, which indicates that hypercapnia induces a buccal activity at lower frequency and at higher rhythm regularity. A wider strip, at high frequency centered at 21 Hz, is present with same characteristics at both values of the pH, corresponding to fast oscillations exhibited on the bursts in S10 signal. For F4, the maps reveal a dark, continuous, low frequency strip (3.7 Hz) at pH=7.8, while the corresponding strip at pH=7.4 is lighter, at 2.5 Hz. Here, hypercapnia induces a lower and more variable frequency of the buccal activity. A second strip is present at both pH values, at about twice the corresponding low frequencies. All these pieces of frequency information are also given in the spectra, which are illustrated for all the signals in Fig. 4. For each premetamorphic tadpole at pH=7.8 and 7.4, the spectra reveal two or three local maxima: the first one occurs systematically in the low frequency range [1.1, 1.6] Hz, and the last one occurs in the high frequency range [20.6, 27.2] Hz. The frequency of the first peak, or buccal low frequency, is slightly lower in hypercapnia than in normocapnia and the amplitude of this peak is higher, but for T5. The high frequency parts of the spectra are very similar at both pH. In the postmetamorphic tadpole, three maxima can be observed in the spectra. In normocapnia, a first peak occurs at low frequency in the range [2.9, 3.7] Hz, systematically followed by the peak at twice this fundamental frequency, i.e. [5.9 7.4] Hz. In two cases (F1, F5) there is a smooth local maximum around 22 Hz. In hypercapnia, the first peak occurs at a lower frequency, in the range [1.5, 2.5] Hz. When present, the second peak is very small in amplitude and a smooth local maximum at high frequency around 22 Hz can 7 be seen in all cases. In contrast with the premetamorphic tadpoles spectra, the amplitude of the first peak is systematically smaller in hypercapnia than in normocapnia, while there is an increase in the high frequency band. The statistical tests performed on the CWT spectral energies in the low frequency (1-10 Hz) and in the high frequency band (20-30 Hz) give the following results. For the effect of metamorphosis, the relative energy in the low frequency band is significantly higher for the postmetamorphic tadpoles at both pH: at pH=7.8, +0.33 (p=0.008) and at pH=7.4, +0.09 (p=0.029), while in the high frequency band it is significantly lower: at pH=7.8, -0.12 (p=0.008) and at pH=7.4, -0.04 (p=0.001). For the effect of the acidosis in premetamorphic tadpoles, there is no significant effect on the relative spectral energy neither in the low frequency band nor in the high frequency band. In postmetamorphic tadpoles, the decrease in the low frequency band is significant: -0.23 (p=0.023), as well as the increase in the high frequency band: +0.07 (p=0.012). 3.2. Amplitude profiles of the buccal canonical cycles In order to visualize in the temporal domain the dynamical modifications induced in the buccal rhythm by hypercapnia, we superimpose the canonical activities at both pH. A graphical comparison is thus obtained by positioning the positive part of the amplitude profile of the canonical activity (i.e. a canonical cycle repeated twice) at pH=7.4 in such a way that the correlation of its first cycle with the first cycle of the canonical activity at pH=7.8 is maximal. These amplitude profiles show that oscillations are systematically present in premetamorphic tadpoles in normo and hypercapnia. For all premetamorphic tadpoles, oscillations occupy the major part of the cycle (Fig. 5, left). These oscillations are rather regular, with an interval of about 35-50 ms between two successive peaks. Under hypercapnia, the cycle duration is systematically increased by about 0.14 s, i.e. by around 20%. This duration increase clearly affects the interburst part of the cycle. The burst amplitude and duration seem very weakly modified. For postmetamorphic tadpoles, the bursts are shorter and exhibit an asymmetry, with a sudden interruption of activity which ends the burst, with very few and small oscillations, if any (Fig. 5, right). Under hypercapnia, the cycle duration is systematically increased by about 0.18 s, i.e. by around 60%. This duration increase clearly affects the interburst part of the cycle also. The burst amplitude is markedly reduced with visible intra-burst oscillations. The N-way ANOVA performed on the mean value of the buccal period (Table 1) indicates that there is no individual effect, but a significant stage effect with a decrease of 335 ms (p=0) and a significant pH effect with an increase of 155 ms (p=0.0004), and no significant cross-effect between pH and stage, which suggests a similar lengthening of the period with the pH for both pre- and postmetamorphic tadpoles. 3.3. Spike contributions to the buccal activity according to their amplitude For T2, the matrices Mk containing the number of spikes in ten successive time bins (i.e. 5 ms) for each of the first four subpopulations (k=1,2,3 and 4) are visualized as maps of gray dots for pH=7.8 and red dots for pH=7.4 (Fig. 6). By definition, the first subpopulation contains the smallest ‘spikes’, with both neuron firings and noise, whose density is higher outside the burst than inside (the burst limits are represented by the green lines in Fig.6): this outside density is greater in hypercapnia than in normocapnia, while the inside burst density is quite the same at both pH values. These features characterize both the number and amplitude profiles, even if this high spike density brings a small contribution to the amplitude profile (Fig. 6). For the second subpopulation, there is a small increase in spike 8 density and amplitude during the burst with slightly smaller values at pH=7.4, particularly during the prolonged interburst part of the cycle. For the third population, there is an increase in spike density and amplitude during the burst, carrying out its plateau part, with a same activity profile of the burst at both pH values. The corresponding spike density has a maximum at 0.31 spikes/5ms = 62 spikes/s. For the fourth subpopulation containing the highest spikes, at both pH values we observe the quasi-absence of spikes outside the burst and an alternation of bands of high and low density of these spikes, resulting to the oscillations of the number and amplitude profiles. In conclusion, in T2 and all premetamorphic tadpoles, in addition to the lengthening of the inactive period, the only visible effect of hypercapnia is the reduction of the second subpopulation spikes during the prolonged interburst. For F2, the matrices Mk (k=1,2,3 and 4) are visualized for pH=7.8 and pH=7.4 in Fig. 7. At pH=7.8, the first subpopulation exhibits a band of very high and constant density at the end of the burst that cannot be of physiological origin: it means that this first peaks/spikes subpopulation is clearly dominated by instrumental noise. For the second population, the spikes are concentrated inside the burst with same densities inside and outside the burst at both pH values, but the duration of the burst is longer at pH=7.4. The spike density of the third subpopulation exhibits an increase during the burst followed by a quasi-absence of spikes after the burst. The spike density inside the burst is smaller in hypercapnia than in normocapnia. These features are accentuated in the fourth subpopulation, where some oscillations are present. For all neurograms from the postmetamorphic tadpoles, as in F2, in addition to the lengthening interburst duration, the main effect of hypercapnia is the strong reduction of the third and fourth subpopulation spikes during the burst. 3.4. Analysis of the lung activity Figure 8 illustrates the intensity of the neuronal activity for the ten neurograms of the postmetamorphic tadpoles represented by the whole corresponding S100 signals, with both buccal and lung components. This figure points out the higher amplitude of the buccal bursts at pH = 7.8 than at pH=7.4. The irregular occurrence of lung bursts is visible in the distribution of the vertical bars indicating these lung bursts. The number of lung bursts is higher at pH=7.4 than at pH=7.8 (but F1, whose number of lung bursts is already high at pH=7.8). The left part of figure 9 illustrates two typical lung bursts observed in the neurograms of the postmetamorphic tadpole F2 at pH=7.8 and pH=7.4 respectively: in this tadpole, all the lung bursts at pH=7.8 are preceded by a small amplitude burst, which is totally absent at pH=7.4. The corresponding mean profiles confirm the presence of these two lung burst types; in addition they show that the mean amplitude of the lung bursts is smaller at pH=7.4 than at pH=7.8. Regarding the contribution to the different spike populations according to their amplitude, the main observation is that the lung burst amplitude is related to the spikes of population V that appears to be specific of the lung activity and seems to be more present in normocapnia than in hypercapnia in F2. Of note, there is no visible temporal structure in the lung bursts as observed in the buccal bursts. 4. DISCUSSION In the present paper, our analysis of the neurograms was done in order to capture some dynamic characteristics of buccal and lung rhythm generator at short-time scale and to evaluate the effects of hypercapnia on these characteristics. In the European frog, Pelophylax ridibundus the major findings of the study are the following: 1. In normocapnia, the frequency of the buccal rhythm in premetamorphic tadpoles is in the range of [75, 96] min-1 and in postmetamorphic tadpoles, in the range [151, 220] min-1, which is higher than all reported values in the American bullfrog (Table 1). 9 2. Fast oscillations in the frequency band 20-30 Hz are present in all tadpoles, revealed by both the spectral analysis and by the buccal amplitude profiles (Fig. 4 and Fig. 5). Such oscillations are visible in other amphibian species under the form of small indentations in the standard filtered neurograms, but they have not been studied. 3. The analysis of spike time positions and amplitudes reveals their respective contributions to the burst structure: the small spikes are uniformly distributed in the burst, while the high ones are present in specific time windows separated by about 35- 50 ms at the origin of the fast oscillations (Fig. 6 and Fig. 7). 4. Metamorphosis induces a significant shortening of the cycle duration of about 335 ms (Table 1) 5. Metamorphosis induces a significant increase of the relative energy in the low frequency band (1-10 Hz) and a significant decrease of relative energy in the high frequency band (20-30 Hz) (Fig. 4). 6. Hypercapnia induces a significant lengthening of about 155 ms of the interburst duration in branchial/buccal rhythm in both premetamorphic and postmetamorphic tadpoles (Table 1 and Fig. 5). 7. In the buccal activity of postmetamorphic tadpoles, hypercapnia induces a significant decrease of the relative energy in the low frequency band (1-10 Hz) and a significant increase of relative energy in the high frequency band (20-30 Hz) (Fig. 4). 8. In the buccal activity of postmetamorphic tadpoles, hypercapnia reduces noticeably in the number of high amplitude spikes, which is related to the low frequency energy decrease (Fig. 7). 9. Lung bursts are present in all postmetamorphic tadpoles, and, as in other species, are more frequent in hypercapnia than in normocapnia (Fig. 8). 10. The lung bursts are characterized by the presence of spikes of high amplitude (population V) From our analysis of the neurograms with their buccal and lung activities, it seems that the lung activity is “phase locked” with the buccal cycles in some preparations. In such preparation, the phase locking may be different depending on the pH (for example see type I or type II lung bursts of figure 9). This dynamic relationship suggests reliable dynamical interconnections between the “lung network” and the “buccal network” of the CPG. Our observations performed on in vitro brainstem preparations of tadpoles of the European frog are similar to those of Taylor et al (2003a and b) with the same type of preparation in Lithobates catesbeianus: hypercapnia is associated with a frequency decrease in buccal oscillations in pre- and post-metamorphic tadpoles through a lengthening of the interburst interval. In the case of premetamorphic tadpoles, the lengthening of the interburst duration induced by hypercapnia leads to a prolonged inactivity of nerve VII motoneurons, which can be attributed to a modification of the buccal rhythm genesis, since there is no lung activity. This hypercapnia-related increase of the interburst in the postmetamorphic tadpoles induces a higher effect on the buccal frequency, since this frequency in normocapnia is much higher in postmetamorphic tadpoles (about 3 Hz) than in premetamorphic tadpoles (about 1.25 Hz). In postmetamorphic tadpoles, this frequency effect is associated to a significant amplitude reduction of the buccal bursts. These two effects that express a decrease in the buccal activity could be the result of a global anesthetic effect of CO2; however, such a global effect would also decrease the lung activity, in contrast with the observed increase of the lung frequency. Both amplitude and frequency reduction of the buccal oscillation are also observed in vivo in adult Lithobates catesbeianus (Kinkead and Milsom, 1994). In air-breathing postmetamorphic tadpoles, several hypotheses have been proposed for the physiological function of the buccal rhythm, which is no longer directly involved in gas exchange. The buccal rhythm ventilates the oropharynx alone: it has been speculated that is optimizes the efficiency of lung breaths 10 by keeping the buccal cavity ventilated with fresh air (Gans et al., 1969). This point of view has been challenged by the attribution of a likely olfactory function to the buccal activity (Foxon, 1964; Vitalis and Shelton, 1990; Jorgensen, 2000). The hypercapnia frequency effect seems to be a buccal Central Pattern Generator (CPG) chemosensitive property already present before metamorphosis and retained after metamorphosis despite the loss of a direct ventilatory function of the buccal activity. Our results evidence that the amplitude decrease of the buccal burst induced by hypercapnia in postmetamorphic tadpoles is due to a smaller number of high amplitude spikes. This reduced motoneuronal activity could result either from a chemosensitivity of a motoneuron population whose threshold may be increased at pH=7.4 or from a less efficient excitatory input from the buccal CPG. In both cases, the motor units that are not recruited in hypercapnia during the buccal activity are likely more efficiently recruited during lung episodes, which occur at a higher frequency than in normocapnia. The CWT spectra of the buccal activity (Fig. 4) reveal the peak at 20-30 Hz (Fig. 4) corresponding to the oscillations observed in S10. These oscillations are mainly visible in the premetamorphic tadpole buccal amplitude profiles (Fig. 5). When focusing on the contribution of the spike sub-populations defined by their amplitude, the spike matrices reveal a quasi-periodic density variation of the fourth population spikes (Fig. 6, top and left), with a period of about 40 ms repeated about 10 times during a buccal burst. This density variation of the highest spikes explains the oscillatory feature of the curves representing their mean number and mean amplitude, and finally the complete amplitude profile (Fig. 5). It is quite unlikely that this temporally organized motoneuron recruitment process is due to intrinsic motoneuron properties, thus we suppose that it is a consequence of the driving of the motoneurons by the buccal CPG, which may send oscillating excitatory inputs to the motoneuronal population. If we suppose that, in premetamorphic tadpoles, the motoneuronal population is quite homogeneous, i.e. weakly differentiated, all these motoneurons are driven in the same way by the oscillating excitatory input and exhibit a higher firing probability at the maxima of the oscillation: the presence of high spikes in the corresponding short time windows may thus result from the summation of individual action potentials occurring in coincidence. In postmetamorphic tadpoles, the peak in the CWT spectra at twice the fundamental frequency [5.9 7.4] Hz is prominent in normocapnia (Fig. 4): it accounts for the specific burst shape, higher and narrower than in the premetamorphic tadpoles (Fig. 3 and Fig. 5), that we relate to the recruitment of a more differentiated motoneuron population. Indeed, the asymmetry of the burst shape exhibited in postmetamorphic tadpole amplitude profiles suggests a progressive recruitment of motoneurons related to size effects of a differentiated population, according to the Henneman size principle (Henneman et al., 1965), where the largest neurons, with the highest threshold are recruitment only at the end of the generator buccal cycle. This progressive recruitment is followed by a sudden stop: the fast and complete de-recruitment of the motoneurons at the end of the burst may be due either to an active inhibition process or to the end of the input burst. In normocapnia, such a recruitment process of the motoneurons may also explain the quasi-absence of fast oscillations at pH=7.8, while present at pH=7.4, although these oscillations are likely structuring the buccal generator activity in both cases. In this hypothetical frame, at pH=7.4, the threshold of CO2 chemosensitive motoneurons may be increased in such a way that the generator activity input is not sufficient to recruit them. However, some small motoneurons, with a low threshold, still respond to this oscillating input (Fig. 4). Fast oscillations have been described in the respiratory neural control system of several mammalian species (for a review see Funk and Parkis, 2002). They are present in inspiratory- related muscles, nerves, and neurons and include both Medium- (MFO) and High-Frequency Oscillations (HFO). The first studies on phrenic and hypoglossal nerve discharges in cats 11 (Richardson and Mitchell, 1982; Cohen et al., 1987; Christakos et al., 1988,1991) revealed frequency peaks in the two bands: 20–50 Hz (MFO) and 60–110 Hz (HFO). Similarly, in adult rats (Marchenko and Rogers, 2006), the time-frequency spectra of phrenic and hypoglossal nerve also exhibited two peaks: one for MFO (37–110 Hz) and one for HFO (156–230 Hz) during eupnea. In mouse (O'Neal et al, .2005), spectral analyses of diaphragm EMG bursts revealed peaks associated with the MFO (20 – 46 Hz) and HFO. In kitten brainstems (Kato et al., 1996), spectra calculated for the inspiratory discharges of facial respiratory nerves showed a stable oscillation in the MFO range only (27-32 Hz) and in newborn rat (Kocsis et al, 1999), this band is also present at (22.8-43.0 Hz). As in kitten, in juvenile rat there is a dominant band in the MFO at 20–55 Hz (Marchenko and Rogers, 2007). Thus, the fast oscillations appear to be dependent on developmental stage, with MFO always present in young animals. The fast oscillations at 20-30 Hz that we observe in Pelophylax ridibundus are reminiscent of MFO in mammals: indeed, Marchenko and Rogers (2007), for instance, attribute the genesis of the MFO band to a central generator that could be of ancestral origin. The presence of fast oscillations in the ventilatory central generator appears as a characteristic that has been retained through the phylogeny, and it would be interesting to look for them in other anuran species, particularly in the well-studied Lithobates catesbeianus. This temporal structure of the buccal cycle, in addition to its fundamental low frequency (about 1 Hz) rhythm, suggests the presence of high frequency pacemaker-like activities of the buccal generator, of cell and/or network origin as observed in several types of locomotion, respiration and mastication CPGs (Harris-Warrick, 2010). We suppose that the activity of the buccal network neurons is the consequence of convergence of two pacemaker inputs with two slightly different frequencies on these neurons. The interference between two such inputs may generate both a slow beating frequency (Lefler Y. et al, 2013) (the buccal fundamental rhythm) and a high frequency related to the mean of the pacemaker frequencies (intraburst oscillations), with or without filter effects (Rose G.J. and Fortune E.S., 1999). In a modeling approach of the buccal and lung CPG, Horcholle-Bossavit and Quenet have proposed (Horcholle-Bossavit and Quenet, 2009) a synchronizing function of the lung part of the CPG in the motoneuron recruitment process, which could account for the high amplitude spikes specific of the lung bursts. The model also implements CO2 sensitivity at the level of the lung part of the CPG to explain the increased lung episodic response to hypercapnia. The results of the present work involve new mechanisms which must be included in the model of the CPG, introducing interactions between pacemakers, in order to implement the buccal CO2 chemosensitivy effects, the fast oscillations structuring the buccal cycle, and the phase- locking of the lung bursts with the buccal activity. 5. CONCLUSION Our results about the buccal ventilatory rhythm control in Pelophylax ridibundus evidence the presence of fast oscillations inside each buccal burst due to coincident firings of motoneurons in quasi-periodic time windows. The effects of hypercapnia on the frequency components, i.e. the low frequency (buccal rhythm) and the high frequency (fast oscillations), suggest that chemosensitivity affects both buccal generator elements and some facial motoneurons properties. ACKNOWLEGMENTS We thank Quentin Michard for technical assistance in the time-frequency analysis and Didier Cassereau and Yacine Oussar for methodological advices. 12 REFERENCES 1. Akay, M., 2005. Hypoxia silences the neural activities in the early phase of the phrenic neurogram of eupnea in the piglet. J. Neuroeng. Rehabil. 30, 32-41. 2. Burggren, W.W., Doyle, M., 1986. Ontogeny of regulation of gill and lung ventilation in the bullfrog, Rana catesbeiana. Respir. Physiol. 66, 279–291 3. Chen, A.K., Hedrick, M.S., 2008. Role of glutamate and substance P in the amphibian respiratory network during development. Resp. Physiol. Neurobiol. 162, 24–31. 4. Christakos, C.N., Cohen, M.I., Barnhardt, R., Shaw, C.F. 1991. Fast rhythms in phrenic motoneuron and nerve discharges. J Neurophysiol. 66, 674-87. 5. Christakos, C.N., Cohen, M.I., See, W.R., Barnhardt, R. 1988. Fast rhythms in the discharges of medullary inspiratory neurons. Brain Res., 463, 362-367. 6. Cohen, M.I., See, W.R., Christakos, C.N., Sica, A.L., 1987. High-frequency and medium-frequency components of different inspiratory nerve and their modification by various inputs. Brain Res. 417, 148-152. 7. Fournier, S. and Kinkead, R., 2008. Noradrenergic modulation of respiratory motor output during tadpole development: role of adrenoceptors. J. Exp. Biol. 209, 3685- 3694. 8. Foxon, G.E.H., 1964. Physiology of the Amphibia, in: Moore, J.A. (Eds.) Blood and respiration. Academic Press, New York, USA, pp. 151–209. 9. Funk, G.D. and Parkis, M.A., 2002. High frequency oscillations in respiratory networks: functionally significant or phenomenological? Resp. Physiol. Neurobiol. 131, 101-20. 10. Galante, R.J., Kubin, L., Fishman, A.P., Pack, A.I., 1996. Role of chloride-mediated inhibition in respiratory rhythmogenesis in an in vitro brainstem of tadpole, Rana catesbeiana. J. Physiol. 492, 545-558. 11. Gans, C., De Jongh, H.J., Farber, J., 1969. Bullfrog (Rana catesbeiana) ventilation: how does the frog breathe? Science 163, 1223-1225. 12. Gdovin, M.J., Torgerson, C.S., Remmers, J.E., 1998. Neurorespiratory pattern of gill and lung ventilation in the decerebrate spontaneously breathing tadpole. Respir. Physiol. 113, 135-146, 13. Gdovin, M.J., Torgerson, C.S., Remmers, J.E., 1999. The fictively breathing tadpole brainstem preparation as a model for the development of respiratory pattern generation and central chemoreception. Comparative Biochemistry and Physiology Part A: Molecular & Integrative Physiology 124, 275–286. 14. Gustafsson, F., 1996. Determining the initial states in forward–backward filtering. IEEE Trans. Signal Process. 44, 988–99. 15. Harris-Warrick R.M., 2010. General Principles of Rhythmogenesis in Central Pattern Networks. Prog Brain Res . 187, 213–222. 16. Hedrick M.S., Chen A.K., Jessop K. L., 2005. Nitric oxide changes its role as a modulator of respiratory motor activity during development in the bullfrog (Rana catesbeiana) Comparative Biochemistry and Physiology Part A: Molecular & Integrative Physiology, 142 (2), 231-240. 17. Henneman, E., Somjen, G. & Carpenter, D. O., 1965. Functional significance of cell size in spinal motoneurons. J. Neurophysiol. 28, 560-580. 18. Horcholle-Bossavit, G., Quenet, B., 2009. Neural model of frog ventilatory rhythmogenesis. Biosystems 97, 35-43. 19. Jorgensen, C.B., 2000. Amphibian respiration and olfaction and their relationships: from Robert Townson (1794) to the present. Biol. Rev. Camb. Philos. Soc. 75, 297- 345. 13 20. Kato, F., Morin-Surun, M.P. and Denavit-Saubie, M., 1996. Coherent inspiratory oscillation of cranial nerve discharges in perfused neonatal cat brainstem in vitro. J. Physiol., 497, 539-549 21. Kinkead, R. and Milsom, W.K., 1994. Chemoreceptors and control of episodic breathing in the bullfrog (Rana catesbeiana). Respir. Physiol. 95, 81–98. 22. Kocsis, B., Gyimesi-Pelczer, K., Vertes, R.P., 1999. Medium-frequency oscillations dominate the inspiratory nerve discharge of anesthetized newborn rats. Brain Res. 818, 180–183. 23. Kogo, N., Perry, S.F., Remmers, J.E. 1994a. Neural organization of the ventilatory activity in the frog, Rana catesbeiana I. J. Neurobiol. 25, 1067-1079. 24. Kogo, N., Remmers, J.E. 1994b. Neural organization of the ventilatory activity in the frog, Rana catesbeina II. J. Neurobiol. 25, 1080-1094. 25. Lefler Y.,Torben-Nielsen B., Yarom Y., 2013. Oscillatory activity, phase differences, and phase resetting in the inferior olivary nucleus. Front Syst Neurosci, 7:22-. 26. Lynn, A., Fuerst, W., 2000. Digital Signal Processing, 2nd ed. John Wiley and Sons, Chichester. 27. Mallat, S., 2000. Une exploration des signaux en ondelettes. Edition de l’Ecole Polytechnique, Paris. 28. Marchenko, V., Rogers, R.F., 2006. Selective loss of high-frequency oscillations in phrenic and hypoglossal activity in the decerebrate rat during gasping. Am. J. Physiol. Regul. Integr. Comp. Physiol. 291,1414-1429. 29. Marchenko, V., Rogers, R.F., 2007. Temperature and state dependence of dynamic phrenic oscillations in the decerebrate juvenile rat. Am. J. Physiol. Regul. Integr. Comp Physiol. 293, R2323-35. 30. Mor, Y. and Lev-Tov, A., 2007. Analysis of Rhythmic Patterns Produced by Spinal Neural Networks. J. Neurophysiol. 98, 2807–2817. 31. O’Neal, I.M.H., Spiegel, E.T., Chon, K.H. and Solomon I.C., 2005. Time-Frequency Representation of Inspiratory Motor Output in Anesthetized C57BL/6 Mice In Vivo. J. Neurophysiol. 93, 1762-1775. 32. Richardson, C.A., Mitchell, R.A. 1982. Power spectral analysis of inspiratory nerve activity in the decerebrate cat. Brain Res., 233, 317-336. 33. Romcy-Pereira, R.N., de Araujo, D.B., Leite, J.P., Garcia-Cairasco, N.A., 2008. Semi- automated algorithm for studying neuronal oscillatory patterns: A wavelet-based time frequency and coherence analysis. J. Neurosci. Methods 167, 384–392. 34. Rose G.J. and Fortune E.S., 1999. Mechanisms for generating temporal filters in the electrosensory system. J Exp Biol. 202: 1281-9. 35. Sakakibara, Y., 1984a. The pattern of respiratory nerve activity in the bullfrog. Jpn J. Physiol. 34, 269-282. 36. Sakakibara, Y., 1984b. Trigeminal nerve activity and buccal pressure as an index of total inspiratory activity in the bullfrog. Jpn J. Physiol. 34, 827-838. 37. Straus, C., Samara, Z., Fiamma, M.N., Bautin, N., Ranohavimparany, A., Le Coz, P., Golmard, J.L., Darré, P., Zelter, M., Poon, C.S., Similowski, T., 2011. Effects of maturation and acidosis on the chaos-like complexity of the neural respiratory output in the isolated brainstem of the tadpole, Rana esculenta. Am. J. Physiol. Regul. Integr. Comp. Physiol. 300, R1163-1174. 38. Straus C., Wilson R.J.A., Remmers J.E., 2000. Developmental disinhibition: turning off inhibition turns on breathing in vertebrates. J Neurobiol 45: 75–83. 39. Taylor, A.C., Köllros, J., 1946. Stages in the normal development of Rana pipiens larvae. Anat. Rec. 94, 7–24. 40. Taylor, B.E., Harris, M.B., Coates, E.L., Gdovin, M.J., Leiter, J.C., 2003a. Central 14 CO2 chemoreception in developing bullfrogs: anomalous response to acetazolamide. J. Appl. Physiol. 94, 1204-1212. 41. Taylor, B.E., Harris, M.B., Leiter, J.C., Gdovin, M.J., 2003b. Ontogeny of central CO2 chemoreception: chemosensitivity in the ventral medulla of developing bullfrogs. Am. J. Physiol. Regul. Integr. Comp. Physiol. 285, R1461–1472. 42. Torgerson, C.S., Gdovin, M.J., Remmers, J.E., 2001a. Sites of respiratory rhythmogenesis during development in the tadpole. Am. J. Physiol. Regul. Integr. Comp. Physiol. 280, R913-920. 43. Torgerson, C.S., Gdovin, M.J., Brandt, R., Remmers, J.E., 2001b Location of central respiratory chemoreceptors in the developing tadpole. Am. J. Physiol. Regul. Integr. Comp. Physiol. 280: R921-928. 44. Torgerson, C.S., Gdovin, M.J., Remmers, J.E., 1997. The ontogeny of central chemoreception during fictive gill and lung ventilation of an in vitro brainstem preparation of Rana catesbeiana. J. Exp. Biol. 299, 2063–2072. 45. Torgerson, C.S., Gdovin, M.J., Remmers, J.E., 1998. Fictive gill and lung ventilation in the pre- and postmetamorphic tadpole brain stem. J Neurophysiol. 80(4), 2015-22. 46. Vialatte, F.B., Martin, C., Dubois ,R., Quenet, B., Gervais, R., Dreyfus, G. 2007. A machine learning approach to the analysis of time-frequency maps, and its application to neural dynamics. Neural Netw. 20, 194-209. 47. Vitalis, T.Z. and Shelton, G.,1990. Breathing in Rana pipiens: the mechanism of ventilation. J. Exp. Biol. 154, 537–556. 48. Wilson, R.J., Vasilakos, K., Harris, M.B., Straus, C., Remmers, J.E., 2002. Evidence that ventilatory rhythmogenesis in the frog involves two distinct neuronal oscillators. J Physiol 540, 557-570. FIGURE LEGENDS Figure 1: Neurogram analyses A) By filtering neurogram Sneur (gray line), two signals are extracted: S100 (black line) and S10 (red line). The neurogram and the signals are illustrated for a premetamorphic tadpole neurogram portion during 1.3 seconds that includes two successive buccal cycles, separated at the level of the local minima of S100 (blue diamonds). Thus, each neurogram is segmented into successive cycles (red dashed vertical lines). B) Identification of the spikes by location of the local maxima on the positive (red stars) and negative (blue stars) parts of the neurogram (gray line, same portion as in A). The green lines define five spike subpopulations (I to V) according to their amplitude Figure 2: Cycle alignment and burst limit definition A) The S10 cycle map is built by positioning and stacking cycles using cross-correlation with a reference cycle (red line). In this map, the cycle amplitude values are coded in gray scale. The blue line represents the proportion of zero padding in the map, between 0 and 1. B) Using the previous positioning of the cycles, spike maps (positive and negative) are built from the spike identification (Fig. 1). The spike amplitude values are represented in gray scale. Positive and negative mean profiles are computed from the respective spike matrices (gray lines); canonical positive and negative cycle amplitude profiles are defined by filtering the previous mean profiles and by cutting these profiles at their most significant portion to define a mean cycle, i.e. where the zero-padding proportion is minimal (between blue lines). The buccal burst is defined as the most active zone of the cycle (between green lines). 15 Figure 3: CWT Time-frequency maps and spectra from S10 signal Upper part: S10 signal during 3 seconds for a premetamorphic tadpole (T2) and a postmetamorphic tadpole (F4) in normocapnia (black line) and hypercapnia (red line) with oscillations on the buccal bursts, particularly visible for T2. Lower part: On the CWT time-frequency maps (shown here for T2 and F4 at pH=7.8 and pH=7.4) of S10 signals with a logarithmic distribution of frequencies in the range [1-100] Hz, normalized spectra in normocapnia (black lines) and hypercapnia (red lines) are superimposed. A maximum of the T2 profiles appears around 20-30 Hz corresponding to the oscillation frequency observed on S10 signal. Figure 4: Effects of hypercapnia on the CWT spectra of premetamorphic and postmetamorphic tadpoles Normalized CWT spectra for premetamorphic tadpoles (T1-T5) and postmetamorphic tadpoles (F1-F5) at pH=7.8 (black lines) and pH=7.4 (red lines). The values of the local maximal frequencies are indicated near the corresponding maxima of the profiles. Figure 5: Effects of hypercapnia on the amplitude profiles of the canonical buccal cycles Amplitude profiles of two canonical buccal cycles computed from the respective positive spike matrices on premetamorphic tadpoles (T1-T5) and postmetamorphic tadpoles (F1-F5) at pH=7.8 (black lines) and pH =7.4 (red lines). The profiles exhibit around 20-30 Hz oscillations as observed in the corresponding S10 signals. The red horizontal segment indicates the interburst period lengthening induced by hypercapnia in both premetamorphic and postmetamorphic tadpoles. Figure 6: Contributions of the spike subpopulations on the buccal burst shape for T2 The left column shows the spike maps at pH=7.8 (black dots) and pH=7.4 (red dots) with the corresponding mean number of spikes for each of the first four subpopulations (I to IV) into bins of 5 ms during a canonical cycle (black line for normocapnia and red line for hypercapnia). The green lines delimit the canonical buccal burst (see Fig. 2). In the right column each amplitude profile is computed from its corresponding spike subpopulation, for pH=7.8 (black line) and pH=7.4 (red line). Figure 7: Contributions of the spike subpopulations on the buccal burst shape for F2 Left and right columns exhibit respectively the number of spikes of each of the first four subpopulations (I to IV) and the corresponding amplitude profiles as in Fig. 6. Note that the number of buccal cycles is lower in hypercapnia than in normocapnia, for the same recording time, due to the marked decrease of the buccal frequency and the higher number of lung episodes which are eliminated from the analysis time. Figure 8: Overview of hypercapnia effect on the buccal and lung activities in neurograms of the postmetamorphic tadpoles (F1-F5). For each postmetamorphic tadpole, the nerve activity recorded during five minutes and filtered to get whole S100 signals, i.e. with both buccal and lung activities, are represented according to a grey level scale. The time position of the lung bursts are indicated by vertical bars, black at pH=7.8 and red at pH=7.4. The grey levels indicate that the buccal activity is higher at pH=7.8 than at pH=7.4. The number of lung bursts is greater at pH=7.4 than at pH=7.8, but for F1. 16 Figure 9: Example of lung bursts Left: typical lung bursts observed in F2 at pH=7.8 (black) and pH=7.4 (red). In F2, all the bursts at pH=7.8 are preceded by a period of small amplitude activity, while such a pre-burst activity is absent at pH=7.4. Right: mean amplitude of the lung burst and the pre-burst activity. Figure 10: Contributions of spike subpopulations II to V on the lung burst shape for F2 Left and right columns exhibit respectively the number of spikes of each of the last four subpopulations (II to V) and the corresponding amplitude profiles as in Fig. 6 and 7. Note that the number of lung bursts is higher in hypercapnia than in normocapnia, for the same recording time. FIGURE 1 500 a.u A B 250 0 (cid:239)250 500 250 0 (cid:239)250 (cid:239)500 FIGURE 2 N 17 V IV III II I 0.25 0.5 0.75 time[s] 1 1.25 1 0 300 250 200 150 100 50 600 500 400 300 200 100 A B 0.25 0.5 0.75 1 0.25 time [s] 0.5 0.75 1 0.25 0.5 FIGURE 3 18 0.05 T3 26 27.2 0.08 F3 1.3 0.04 T4 1.3 2.4 26 23.6 0.08 F4 3.7 2.5 7.4 5.1 20.6 1.6 1.2 2.3 T5 27.2 26 3.5 F5 7.1 2.2 20.6 22.6 4.4 6.4 26 1.5 1 10 100 1 10 100 1 10 100 Hz FIGURE 4 a.u. 1.3 0.06 1.2 0.06 T1 0.06 T2 1.1 1.3 1.4 2.5 2.9 22.6 26 0.06 0.1 F1 2 2.9 2 1.5 20.6 3.7 3.1 21.5 0.1 F2 2.4 2.2 3.2 5.9 21.5 4.2 21.5 1 10 100 1 6.1 23.6 20.6 100 4.2 10 FIGURE 5 a.u 150 0 300 0 150 0 150 0 100 0 0.5 19 F1 0.12 s F2 0.18 s F3 0.32 F4 0.13 s F5 0.16 s 0.75 0.25 time [s] 200 a.u 0.08 s T1 0.22 s T2 0.1 s T3 T4 T5 0.14 s 0.15 s 1 0 300 0 400 0 400 0 600 0 1.5 time [s] FIGURE 6 Mean number of spikes per cycle Mean number of spikes per cycle 0.8 Amplitude a.u. Amplitude a.u. 0.4 0.8 0.4 0.8 0.4 1.6 0.8 IV III II I 100 50 0 100 50 0 100 50 0 100 50 0 0.25 0.5 time [s] 0 0 0.75 0.25 0.5 time [s] 0.75 FIGURE 7 Mean number of spikes per cycle Mean number of spikes per cycle Amplitude a.u. Amplitude a.u. 20 0.4 0.2 0.4 0.2 0.4 0.2 1.6 0.8 IV IV IIIIII IIII I I 100 50 0 100 50 0 100 50 0 100 50 0.3 0 0 0.4 0.1 0.2 time [s] 0.3 0.4 0 0.1 0.2 time [s] FIGURE 8 21 FIGURE 9 Amplitude a.u. 1000 0 (cid:239)1000 1000 0 Amplitude a.u. (cid:239)1000 0 0.25 time [s] 0.5 900 Amplitude a.u. 600 300 0 900 Amplitude a.u. 600 300 0 0 0.25 time [s] FIGURE 10 Mean number of spikes per burst Mean number of spikes per burst Amplitude a.u. Amplitude a.u. 0.5 1 0.5 1 0.5 1 0.5 1 0.5 0 V V IVIV IIIIII II II 600 400 200 0 600 400 200 0 600 400 200 0 600 400 200 0.25 time [s] 0.5 0 0 0.25 time [s] 0.5
1805.09001
6
1805
2019-04-24T03:07:16
One-to-one Mapping between Stimulus and Neural State: Memory and Classification
[ "q-bio.NC", "cs.LG", "stat.ML" ]
Synaptic strength can be seen as probability to propagate impulse, and according to synaptic plasticity, function could exist from propagation activity to synaptic strength. If the function satisfies constraints such as continuity and monotonicity, neural network under external stimulus will always go to fixed point, and there could be one-to-one mapping between external stimulus and synaptic strength at fixed point. In other words, neural network "memorizes" external stimulus in its synapses. A biological classifier is proposed to utilize this mapping.
q-bio.NC
q-bio
One-to-one mapping between stimulus and neural state: Memory and classification Sizhong Lan1 1China Mobile Research Institute, Beijing, 100053, Chinaa) (Dated: 25 April 2019) Synaptic strength can be seen as probability to propagate impulse, and according to synaptic plasticity, function could exist from propagation activity to synaptic strength. If the function satisfies constraints such as continuity and mono- tonicity, the neural network under external stimulus will always go to fixed point, and there could be one-to-one map- ping between the external stimulus and the synaptic strength at fixed point. In other words, neural network "memorizes" external stimulus in its synapses. A biological classifier is proposed to utilize this mapping. 9 1 0 2 r p A 4 2 ] . C N o i b - q [ 6 v 1 0 0 9 0 . 5 0 8 1 : v i X r a I. INTRODUCTION Known experiment results show that synaptic connection strengthens or weakens over time in response to increases or decreases in impulse propagation1. It is also postulated that "neurons that fire together wire together"2,3. This biochem- ical mechanism, called synaptic plasticity4,5, is believed to play a critical role in the memory formation6 -- 9, although it is still argued if synapse is the sole locus of learning and memory10,11. Meanwhile, a synapse propagates impulses stochastically12 -- 14, which means that synaptic strength could be measured with the probability of propagating an impulse successfully. With this probabilistic treatment we find out that, in the plasticity process a synapse's strength would be inevitably attracted towards the same fixed point regardless of its initial strength, and for a neural network there could ex- ist a one-to-one mapping between the external stimulus from environment and the synapses' strength at fixed point. This one-to-one mapping serves the very purpose of ideal mem- ory: to develop different stable neural state for different stim- ulus from the environment, and develop the same stable neural state for the same stimulus no matter what state is initialized with. It follows that the synapses alone could sufficiently give rise to persistent memory: they could the sole locus of learn- ing and memory. The remainder of paper goes as follows. Section II iden- tifies the constraints under which synaptic plasticity of one synaptic connection leads to its fixed state and one-to-one stimulus-state mapping (memory). Section III extends the concepts of fixed state and one-to-one mapping for the neu- ral network consisting of many synaptic connections. Section IV proposes a simple neural classifier utilizing this memory to classify handwritten digit images. II. SYNAPTIC CONNECTION AND ITS FIXED POINT Let us start with one synaptic connection as shown in FIG 1. In nature, synapses are known to be plastic, low-precision and unreliable15. This stochasticity allows us to assume synap- tic strength s to be the probability (reliability) of propagat- a)Electronic mail: [email protected] FIG. 1. A synaptic connection with strength s is directed from neuron 1 to neuron 2. The stimulus from environment or upstream neurons stimulates neuron 1 to fire action potential with probability x. The synaptic connection propagates nerve impulse (action potential) to neuron 2. As a result, neuron 2 is stimulated to fire with probability y. That is, neuron 1 and 2 fire simultaneously ("fire together") with probability y. ing a nerve impulse through, instead of being weight (usu- ally unbounded real number) as in Artificial Neural Network16 (ANN). Easily we have y=xs where x,s,y∈[0,1]. Now we treat synaptic plasticity, i.e. the relation between synaptic strength s and simultaneous firing probability y, as a function s∗ = λ (y). (1) Here s∗∈[0,1] represents the target value that a connection's strength will be strengthened or weakened to if the connec- tion is under constant simultaneous firing probability y (while s in y=xs represents current strength). By y=xs and Eq. (1), we have s∗=λ (xs) stating that, under constant stimulus prob- ability x, the connection initialized with strength s will evolve towards s∗. Function λ of Eq. (1) truly links "firing together" and "wiring together". For comparison, Hebbian learning rule2 treats synaptic plasticity, in the context of ANN, as a function ∆w=ηxy to learn connections' weight from the training pat- terns; the function translates "firing together" into "neuron's input and output both being positive or negative". Different from ANN, our model actually makes no assumption of neu- ron being computational unit, and aims to show that with λ stimulus could sufficiently and precisely control the enduring fixed state of synaptic connection. The following reasoning will hinge on this "target strength function" λ , and we will put constrains on this uncharted function to see how they af- fect the dynamics of connection strength and most importantly how stimulus is one-to-one mapped to the strength at fixated state. 2 FIG. 3. The simulation results of Algorithm 1 for four typical λ func- tions. For each λ , eleven trails parameteried with incremental initial strength s0 are run for 105 iterations; all trails share the same stim- ulus probability x=0.8. In each λ 's subfigure, the left chart depicts λ (y) as black line, its horizontally scaled λ (xs)=λ (0.8s) in red line and fixed points as blue dots; the right chart shows the strength tra- jectories starting from incremental s0. (a) λ (y)=0.9y+0.05. There exists one single fixed point for λ (0.8s). All strength trajectories converge to this fixed point. (b) λ (y)=0.5sin(4πy)+0.5. There are three fixed points for λ (0.8s), two of which are stable ones for the (c) λ (y)=−y+1. All trajectories to converge to with oscillation. trajectories converge to one single fixed point. (d) λ (y) is discontin- uous at y=0.5. There is no fixed point since there is no s∈[0,1] such that λ (0.8s)=s, and consequently the trajectories don't converge. In- stead, they fluctuate within "fixed interval", which as we will see is a useable compromise of fixed point. strengthened or weakened. Here the crucial Brouwer's theo- rem is a fixed-point theorem in topology, which states that, for any continuous function f (t) mapping a compact convex set (e.g. interval [0,1] in our case; could be multi-dimensional) to itself, there is always a point t+ such that f (t+)=t+. More- over, as illustrated in FIG 2, given any initial value the strength is always attracted towards fixed point. Therefore, a gentle constraint of continuity on λ function could preferably drive synaptic connection to the fixed state. To verify connection strength's tendency towards fixed points, we design Algorithm 1 to simulate our connection model. In this simulation18, recent simultaneous firings are recorded and the rate is supposed to approximate the simulta- neous firing probability y; the connection updates its strength FIG. 2. Two examples of λ (xs) are depicted as red bold lines, and their fixed points as blue dots. (a) Given any initial s1<λ (xs1), there must exist a fixed point s+∈(s1,1]; strength s tends to increase from s1 as long as target strength λ (xs)>s. Given any initial s2>λ (xs2), there must exist a fixed point s+∈[0,s2); strength s tends to de- crease from s2 as long as target strength λ (xs)<s. Controlled by these two tendencies, s will reach and stay at fixed point s+ such that s+=λ (xs+). (b) There are three fixed points s+ 2 ,s+ 3 . Start- ing from any initial s1∈(s+ 1 . Starting from any initial s2∈(s+ 3 ,s+ 2 . Then strength tends to leave unstable fixed point s+ 1 or s+ 2 . Note that if countable fixed points exist for λ (xs), one of them must be stable. 1 ,s+ 2 ), strength increases to s+ 3 for stable s+ 1 ,s+ 3 ), strength decreases to s+ Algorithm 1 connection strength's tendency to fixed points Input: stimulus probability x, initial synaptic strength s0, target strength function λ , strength step ∆s, and interations I. neuron 1 and 2 fire together: recorder[p]←1. endif if recorder has been traversed once (i≥104) then preset current pointed entry of recorder: recorder[p]←0. pick random r1 and r2 from uniform distribution Uni f (0,1). if x>r1 and s>r2 then Output: trajectory of strength s. 1: initialize fire-together recorder (104-entries array): recorder⇐0. 2: initialize fire-together recorder pointer: p←0. 3: initialize current strength: s←s0. 4: for i=0 to I do 5: 6: 7: 8: 9: 10: 11: 12: 13: 14: 15: 16: 17: 18: 19: 20: 21: end for end if forward recorder pointer: p←(p+1) mod 104. set y with the proportion of 1-entries in recorder. set target strength: s∗←λ (y). if s∗>s then step-increase current strength: s←min(s+∆s,1). step-decrease current strength: s←max(0,s−∆s). end if if s∗<s then end if Here is our first constraint: λ is continuous on y. This constraint is neurobiologically justifiable regarding synaptic plasticity, since sufficiently small change in impulse probabil- ity would most probably result in arbitrarily small change in synaptic strength. In that case, given any x, λ (xs) is a continu- ous function on s from unit interval [0,1] to unit interval [0,1], and according to Brouwer's fixed-point theorem17 there must exist a fixed point s+∈[0,1] such that s+=λ (xs+): connection strength at s+ will evolve to s+ and hence fixate, no longer by a small step ∆s=10−4 each iteration to the direction of tar- get strength. As shown in FIG 3, we run the simulation for four typical target strength functions, and the strength trajec- tories resulted show that the constraint of continuity ensures the tendency towards fixed points given any initial strength. Our goal is to establish a one-to-one mapping between the stimulus and the connection strength at fixed point. Specif- ically, we could (1) given any stimulus x∈[0,1], identify the fixed point s+ of connection strength without ambiguity; (2) given any strength s+∈[0,1] at fixed point, identify stimulus x without ambiguity. Among the four target strength func- tions in FIG 3, λ (y)=0.9y+0.05 and λ (y)=−y+1 can lead to one-to-one stimulus-strength mapping. Given any stimulus x, a synaptic connection equipped with one of these functions will have one single fixed point of strength regardless of its initial strength, such that the relation between stimulus and fixed point strength can be treated as a function s+=θ (x). In FIG 4, simulation shows that θ could be strictly monotonic and hence one-to-one mapping from x to s+, such that θ (x) has one-to-one inverse function θ−1(s+). By contrast, FIG 5 shows that λ (y)=0.5sin(4πy)+0.5 cannot ensure the unique- ness of fixed point and thus there is no such one-to-one θ (x); FIG 6 shows that there is no θ either for the discontinuous λ function in FIG 3(d). 3 5. simulation results The of Algorithm 1 FIG. for λ (y)=0.5sin(4πy)+0.5 in FIG 3(b) to identify the relation be- (a) Given x=0.3 there is one single fixed point tween x and s+. (b) As with x=0.8 in FIG 3(b), regardless of initial strength s0. given x=0.9 there are two stable fixed points. Higher s0 converges to higher fixed point; lower s0 converges to lower one; No convergence (c) Trails with incremental x to the middle unstable fixed point. are run and the averaged s+ values are depicted as in FIG 4. From x=0.65 and upwards, there are two possible stable fixed points to converge to depending on what value initial strength is randomized to, which means that there exists no θ function from x to s+. FIG. 4. The simulation results of Algorithm 1 to reveal the relation of stimulus probability x and fixed point strength s+. For each λ , sim- ulation is parameterized with incremental x (rather than x=0.8 as in FiG 3) and randomized s0. Ten trails are run for each incremental x, and the ten converged s values are averaged to be the s+ value corre- sponding to its input x. The red line represents the averaged s+ values from simulation, while the blue line represents the true s+=θ (x). (a) Simulation is parameterized with λ (y)=0.9y+0.05 and the results match θ (x)=0.05/(1−0.9x) which is monotonically increasing. (b) Simulation is parameterized with λ (y)=−y+1 and the results match θ (x)=1/(1+x) which is monotonically decreasing. In fact, we can pinpoint more constraints on λ as the con- ditions for function θ to be one-to-one mapping. In addi- tion to constraint of continuity, let λ(y) be strictly mono- tonic on [0, 1] and hence one-to-one; let λ(0)(cid:54)=0 to rule out fixed point s+=0. In that case, λ has inverse function λ−1(s) which is strictly monotonic between λ (0) and λ (1), and given any fixed point strength s+ between we can identify stimu- lus x=λ−1(s+)/s+. That is, function θ−1(s+) = λ−1(s+)/s+ exists. Let λ−1(s)/s be strictly monotonic between λ(0) and λ(1). Then given any stimulus x∈[0,1] there is one sin- gle fixed point s+ such that x=λ−1(s+)/s+. That is, function s+=θ (x) exists. Both of λ (y)=0.9y+0.05 and λ (y)=−y+1 FIG. 6. The simulation results to find the θ function with respect to the discontinuous λ in FIG 3(d). When x(cid:38)0.6, strength can evolve to any point within a "fixed interval" each time simulation is finished. The absence of fixed point doesn't allow the existence of θ. obey all those constraints and their one-to-one θ functions can be verified by the simulation results in FIG 4, whereas λ (y)=0.5sin(4πy)+0.5 is not even strictly monotonic. How- ever, neither λ (y)=0.9y+0.05 nor λ (y)=−y+1 is ideal for our purpose. Guided by these constraints, we choose λ func- tion carefully such that its derived θ (x) function is monoton- ically increasing and range of which spans nearly the entire [0,1] interval, as shown in FIG 7. Of all the λ constraints, continuity and strong monotonicity are reasonable require- ments of consistency on the neurobiological process of synap- tic plasticity, whereas λ (0)(cid:54)=0 and strong monotonicity of λ−1(s)/s are rather specific and peculiar claims. Admittedly, those λ constraints need to be supported by experimental evi- dences. Now we have the one-to-one (continuous and strictly mono- tonic) functions λ , λ−1, θ and θ−1, and in those functions ini- tial strength s0 is irrelevant. Given s+ we can identify x and y without ambiguity, and vice versa. Our interpretation of these mappings is, the synaptic connection at fixed point precisely "memorizes" the information of what (stimulus) it senses and 4 Stimulus and strength uniquely determine impulses prop- agation within neural network, so there exists a mapping Ψ:(X,S)→Y . Presumably, the mapping Ψ is continuous on (1), there exists a mapping Λ:Y→S∗ such that S. By Eq. s∗ i j=λi j(yi j) for each yi j in Y and its counterpart s∗ i j in S∗. Here S∗∈[0,1]c is c-dimensional vector of connections' tar- get strength, and mapping Λ could be visualized as a vec- tor of target strength functions such that entry Λi j is λi j. Then with mapping Ψ and Λ we have a composite mapping Λ◦Ψ:(X,S)→S∗. If each λi j function is continuous on its yi j, mapping Λ◦Ψ must be continuous on S and according to Brouwer's fixed-point theorem given X there must exist one fixed point S+∈[0,1]c such that Λ◦Ψ(X,S+)=S+. And under constant stimulus X, neural network will go to fixed point S+ as each connection i(cid:32) j goes to its fixed point s+ i j. Our simu- lation verifies that tendency as shown in FIG 9. In this simula- tion, impulses traverse the neural network stochastically such that each neuron is fired at most once per iteration; synaptic connections update their strength as in Algorithm 1. √ FIG. 7. (a) Our choices of λ functions are λL(y)=0.99 y+0.01 in 1+e−4.4(y+0.01) −1 in green. Here λT is a segment of blue and λT (y)= shifted and scaled Sigmoid function. They both obey the λ con- straints as discussed previously. (b) Simulation results show that, λL leads to linear-like θL in blue such that θL(x)≈x, and λT leads to threshold-like θT in green. 2 how it responses (with impulse propagation). III. NEURAL NETWORK AND ITS FIXED POINT FIG. 8. A neural network (of one or multiple agents) consists of n≥2 neurons and c≥1 directed synaptic connections. An example of n=8 and c=7 is depicted. Each neuron receives stimulus from the envi- ronment with probability and propagates out nerve impulses through- out the synaptic connections, e.g., triggered by stimulus neuron 1 propagates impulses stochastically down along the directed paths 1(cid:32)7(cid:32)8(cid:32)2 and 1(cid:32)7(cid:32)8(cid:32)5(cid:32)3. Cyclic path (e.g. 3(cid:32)8(cid:32)5(cid:32)3) is allowed and yet loop (e.g. 3(cid:32)3) isn't. Each neuron could have either outbound or inbound connections, or neither, or both. Now let us turn to the neural network shown in FIG 8. A neural network could be treated as an "aggregate connection" as it turns out. We shall see that, the definitions and reasoning for neural network align well with neural connection in last section. As with synaptic connection, we can describe a neural net- work by defining (1) the external stimulus as an n-dimensional vector X∈[0,1]n in which each xi is the probability of neuron i receiving stimulus; (2) the strength of all connections as a c- dimensional vector S∈[0,1]c in which each si j is the strength of connection from neuron i to neuron j (denoted as i(cid:32) j); (3) the simultaneous firing probabilities over all connections as a c-dimensional vector Y∈[0,1]c in which each yi j is the simul- taneous firing probability over i(cid:32) j. In fact, one single neural connection is a special case of neural network with c=1 and n=2. FIG. 9. Simulation results of neural network's tendency for the four typical λ functions as in FIG 3. The neural network has n=8 and c=19. The following observations hold true for any external stimu- lus and connections configuration: (a) If all connections are equipped with λ (y)=0.9y+0.05, the whole neural network has one single fixed point and the trajectories of mean of all connections' strength con- verge to one point. (b) λ (y)=0.5sin(4πy)+0.5. Because each con- nection has two stable fixed points, there are 219 stable fixed points for the whole neural network and 20 possible convergence points of strength mean. (c) λ (y)=−y+1. There is one single fixed point for the neural network. The trajectories converge to one point. (d) Dis- continuous λ . The neural network has no fixed point as each synaptic connection has no fixed point. The trajectories don't converge. Generally the number of stable fixed points for a neural net- work is ∏c fi j where each fi j is the number of stable fixed points of i(cid:32) j. As in FIG 9(b), ∏c fi j can be enormous when each fi j ≥ 2. As with synaptic connection, our goal is to estab- lish one-to-one mapping between stimulus X and fixed point S+ for neural network and meanwhile keep initial strength S0 out of picture. λ 's continuity alone cannot ensure the unique- ness of fixed point, such that S0 can determine which fixed point to go for. Now with all the λ constraints, we have: (1) Λ is a one-to-one mapping and thus has inverse mapping Λ−1:S∗→Y ; (2) there exists a mapping Θ:X→S+, because un- der stimulus X the neural network will go to the same unique fixed point S+ no matter what initial strength S0 to begin with; (3) if Θ is a one-to-one mapping, Θ has inverse mapping Θ−1:S+→X. With mapping Λ, Λ−1, Θ and Θ−1 being one- to-one, given S+ we can identify X and Y without ambiguity, and vice versa. Therefore, the same interpretation with respect to synaptic connection could apply here: the neural network at fixed point precisely "memorizes" the information about the stimulus on many neurons and the impulse propagation across many connections. Nevertheless, even all of λ constraints are not sufficient to secure one-to-one Θ:X→S+ for a neural network, as opposed to the neural connection. Here is a case. For Θ to be one- to-one, all neurons must have outbound connection. Other- wise, e.g., for a neural network with three neurons (say 0, 1 and 2) and two connections (say 0(cid:32)1 and 1(cid:32)2), stimu- lus X1=(1,1,0) and X2=(1,1,1) will result in the same fixed point because stimulus on neuron 3, no matter what it is, af- fects no connection. Or equivalently, for Θ to be one-to-one, the definition of X should consider only the neurons with out- bound connections such that X's dimension dim(X)≤n. In the perspective of information theory19, many-to-one Θ in- troduces equivocation to the neural network at fixed point, as if information loss occurred due to noisy channel. If dim(X)>dim(S)=c, mapping Θ conducts "dimension reduc- tion" on stimulus X, and information loss is bound to occur. Here is a trivial case regarding stimulus dependence. Con- sider a neural network with 0(cid:32)2, 1(cid:32)2 and 2(cid:32)3, and stim- ulus X=(x0,x1). When the neural network is at fixed point, 12x0 Pr(10) where Pr(10) is the proba- x2=x0s+ bility of neuron 1 being stimulated conditional on neuron 0 being stimulated. Pr(10)(cid:54)=x1 if stimulus on neuron 1 and 2 are not independent. Pr(10) affects s+ 23 and hence S+, or in other words the neural network at fixed point gains the hidden information of Pr(10). However, if Pr(10) varies, given mere X there will be uncertainty about S+ such that mapping Θ doesn't exist unless stimulus X is "augmented" to X=(x0,x1,Pr(10)). 12−s+ 02+x1s+ 02s+ 5 its particular training stimulus, and then a testing stimulus is tested on all g neural networks independently to see which gets the most connections propagated. For simplicity we as- sume testing itself doesn't jeopardize the fixed points of neural networks. And most importantly we assume that for each neu- ral network given any stimulus there is one single fixed point such that mapping Θ:X→S+ exists. Consider a neural network in the classifier to be trained by X to fixed point S+ and then tested by X. In other words, neural network memorizing X as S+ is tested by X. Because impulses propagate across the neural network stochastically, the count of synaptic connections propagated in one test should be random variable. Let it be Z XX. Then for the neural network in FIG 8 Z XX =∑c zi j where each r.v. zi j∼Bernoulli(xis+ i j ), i.e., synaptic connection i(cid:32) j is propagated with probability xis+ in the test such that i j i j. Easily zi j's expected value is E[zi j]=xis+ Pr(zi j=1)=xis+ i j, and its variance is Var(zi j)=xis+ i j ). By central limit theorem, Z's distribution could tend towards Gaussian-like (bell curve) as c increases, even if all zi j are not independent and identically distributed. We have i j (1−xis+ E[Z XX ] = c ∑ i(cid:32) j E[zi j] = c ∑ i(cid:32) j xis+ i j . (2) And when c is large, Var(Z XX ) ≈ c ∑ i(cid:32) j Var(zi j) = c ∑ i(cid:32) j i j (1− xis+ i j ). xis+ (3) For any i(cid:32) j, in the training stage because S+=Θ( X) we i j =θi j( X), and in the testing stage xi is uniquely deter- have s+ mined by S+ and X such that xi is a function of X and X. IV. AN APPLICATION FOR CLASSIFICATION Ideally, a neural network with memory of stimulus X -- formally, mapping Θ casts memory of stimulus X as fixed point S+ -- should response to stimulus X more "intensely" than the neural network with different memory responses to X. Memory would manifest itself as impulses propagation throughout ensemble of neurons2,20 -- 22. Thus, it is natural to differentiate response by counting the neurons fired or synap- tic connections propagated by impulses. Given the reasoning that synapse could be the sole locus of memory10,11, we adopt the count of synaptic connections propagated as a macro- scopic measure of how intensely memory responses to stim- ulus or stimulus "recalls" memory. And accordingly we pro- pose a classifier consisting of g neural networks, which clas- sifies stimulus into one of g classes by the decision criteria of which neural network gets the most synaptic connections propagated. Reminiscent of supervised learning23, each neu- ral network of our classifier is trained to its fixed point by FIG. 10. The depicted neural network is basically the general one in FIG 8 except that, to mimic real-life nervous system, an array of sensor neurons are specialized for receiving stimulus from no other neurons but the environment. There are 64 sensor neurons to accom- modate 8×8-pixel image, and the rest are a cluster of 50 neurons. Each sensor neuron has 6 outbound connections towards cluster, and each cluster neuron has 5 outbound connections towards within clus- ter. Connections are randomly put between neurons before training. We experiment with this classifier to classify handwritten digit images24. Ten identical neural networks (hence g=10) of FIG 10, each designated for a digit from 0 to 9, are trained to their fixed points by their training images in FIG 11 as stimulus, and then testing images, also as stimulus, are clas- sified into the digit whose designated neural network gets the FIG. 11. A digit image has 8×8=64 pixels, and pixel grayscale is normalized to the value between 0 and 1 (by dividing 16) as stimulus probability. The upper row shows samples of digit images, and the lower row shows the better written "average images", each of which is actually pixel-wise average of a set of images of a digit. Each neu- ral network is trained in each iteration by the same "average image", or equivalently in each iteration by image randomly drawn from the set of images. biggest Z value. We run many tests to evaluate classification accuracy, and collect Z values to approximate r.v. Z's dis- tribution. With all synaptic connections equipped with λL in FIG 7, the classifier has accuracy ∼44%, and ∼51% with λT . Note that, equipped with λL or λT , the neural network of FIG 10 will have one-to-one ΘL or ΘT according to last section. FIG 12 and FIG 13 show that, in positive testing (e.g. digit-6 image is tested in neural network trained by digit-6 images), Z's expected value (sample mean) could be considerably big- ger than that in negative testing (e.g. digit-6 image is tested in neural network trained by digit-1 images), so as to discrim- inate digit-6 images from the others. Given the same testing image classification target can be different test by test since the ten Z outcomes are randomized. To improve classifica- tion accuracy, we shall distance the distribution of positive testing Z as far as possible from those of negative testing Z. We present another two special neural networks in FIG 14 to demonstrate how our classifier utilizes memory to classify im- ages and how to improve its accuracy in the neurobiological way. FIG. 12. The histogram (in probability density form) of Z. To collect Z values, a digit-6 image is tested many times on each of the ten trained neural networks. All connections are equipped with λT . Z66 of positive testing is in red, and the other nine Zk6 of negative testing, where k=0,1,2,3,4,5,7,8,9, are in gray. Z's sample mean for each digit is depicted as vertical dotted line. When the classifier adopts ten neural networks of FIG 14(a) and equips all connections with λL in FIG 7, classification ac- curacy is ∼31% and Z's distribution for testing digit-6 im- ages is shown in FIG 15(a). We already know that λL makes θL(x)≈x. Then for one test we have E[Z XX ] = 64 ∑xiθi( xi) = 64 ∑xiθL( xi) ≈ 64 ∑xi xi = X(cid:124)X. (4) 6 FIG. 13. The histograms of Z for all ten digits. For each digit, ran- domly drawn testing image, instead of the same one, is used in each test. From digit 0 to 9, classification accuracy is approximately 70%, 41%, 56%, 42%, 53%, 33%, 77%, 51%, 57% and 32%. Generally, better Z-distribution separation of positive and nagative testing re- sults in higher classification accuracy. (2) and Eq. i =θi( xi). By Eq. FIG. 14. These two neural networks inherit the sensor-cluster struc- ture of FIG 10. (a) Each sensor neuron connects to one single clus- ter neuron such that each pixel stimulus xi only affects one single connection. Then s+ (3), we have E[Z XX ]=∑64 xiθi( xi) and Var(Z XX )≈∑64 xiθi( xi)[1 − xiθi( xi)]. (b) Each sensor neuron connects to its own dedicated cluster of many neurons and synaptic connections, and the clusters are of different sizes. In that case, in a test each xi causes ωi (instead of just one) synaptic connections to be propagated with probability xis+ i or none with probability 1−xis+ i . When each ωi is a nonrandom variable, we have E[Z XX ]=∑64 xiθi( xi)ωi and Var(Z XX )≈∑64 xiθi( xi)[1 − xiθi( xi)]ω2 i . Here X(cid:124)X is the dot product of training vector X∈[0,1]64 and testing vector X∈[0,1]64. Generally, the dot product of two vectors, a scalar value, is essentially a measure of simi- larity between the vectors. The bigger E[Z XX ] is, the more in- tensely neural network with memory of training X responses to testing X, and the more similar X and X are to each other. Therefore, Eq. (4) simply links otherwise unrelated neural re- sponse intensity and stimulus similarity. By comparing ten E[Z XX ] values, we can tell which X is the most similar to X and hence which digit is classification target. Only, Z XX value from test actually deviates around the true E[Z XX ] randomly, which makes it a useable and yet unreliable classification cri- teria. When the classifier equips all connections with threshold- like λT in FIG 7, classification accuracy raises to ∼44%. By comparing FIG 15(b) with FIG 15(a), the distance between Z66's distribution and the other nine Zk6,k(cid:54)=6's distribution is bigger with threshold-like λT than with linear-like λL. This accuracy improvement can be explained conveniently with a true threshold function (or step function) (cid:40) θstep(x) = 0, 0 ≤ x < xstep xstep ≤ x ≤ 1 1, . Of the sum terms in ∑64 xi xi of Eq. (4), θstep basically di- minishes small xi∈[0,xstep) to 0 and enhances big xi∈[xstep,1] to 1, such that most probably E[Z66] would increase by hav- ing xi=1 in the sum terms with big xi while the other nine E[Zk6,k(cid:54)=6] would decrease by having xi=0 in the sum terms with big xi, so as to preferably increase E[Z66]−E[Zk6,k(cid:54)=6]. And likewise Var(Z) would most probably decrease. As a re- sult, θstep increases the distance between the distribution of Z66 and Zk6,k(cid:54)=6 and thus better separates them. FIG. 15. The Z histogram of testing digit-6 images for three dif- ferent classifier settings. For each classifier setting, values of Z66 and Zk6,k(cid:54)=6 are transformed to z-scores (i.e. the number of standard deviations from the mean a data point is) with respect to the distri- bution of all Zk6,k(cid:54)=6's values combined. The distance between the distribution of Z66 and Zk6,k(cid:54)=6 is approximately evaluated by E[Z66] and standard deviation σ (Z66). (a) Classifier with neural networks of FIG 14(a) and λL. E[Z66]≈0.93 and σ (Z66)≈0.98. (b) Classi- fier with neural networks of FIG 14(a) and λT . E[Z66]≈1.33 and σ (Z66)≈0.9. (c) Classifier with neural networks of FIG 14(b), λL and ωi( xi)=100 xi 3. E[Z66]≈1.39 and σ (Z66)≈0.87. FIG 14(b) provides another type of neural network to im- prove classification accuracy without adopting threshold-like λ function for all synaptic connections. Let the linear-like 3 simply for example. λL be equipped back and take ωi=100 xi With this setting our classifier has accuracy ∼47%. Here we 7 TABLE I. Classification accuracy on different classifier settings. In each setting, neural network can be the one illustrated in FIG 10, FIG 14(a) or 14(b), and target strength function can be one of those in FIG 3 and FIG 7. The accuracy listed is the average of many out- comes taken from the same trained classifier, and thus could fluctuate slightly from one training to another. λ or θ functions λL(y)=0.99 λT (y)= θstep λ (y)=0.9y+0.05 λ (y)=0.5sin(4πy)+0.5 λ (y)=−y+1 Discontinuous λ FIG 14(b) 47% 51% 60% b 19% 2% 1%d 28% FIG 14(a) 31% 44% 48%a 16% 6% 5% 20% FIG 10 44% 51% - 14% 5%c 4% 23% √ y+0.01 1+e−4.4(y+0.01) −1 2 a xstep is set to 0.6. b xstep is set to 0.2. c Accuracy under 10% is actually worse than wild guessing. d If classification criteria is changed to "which neural network gets the fewest synaptic connections propagated", the accuracy will be ∼40%. 4) where 100 xi 0.1) (here 4√ have E[Z XX ]≈∑64 xi(100 xi transforms xi∈[0, 4√ and transforms xi∈[ 4√ 4, like θstep, actually 0.1≈0.56) to within [0,10) 0.1,1] to across [10,100] -- again, the strong training pixel-stimulus are greatly weighted while the weak ones are relatively suppressed. As shown in FIG 15(c) the distance between the distribution of Z66 and Zk6,k(cid:54)=6 is in- creased compared to FIG 15(a). Here our neurobiological in- 3 is, the training stimulus af- terpretation regarding ωi=100 xi fects not only synaptic strength, but also the growth of neuron cluster in the replication of neuron cells and in the formation of new synaptic connections. Again this claim needs to be supported by evidences. TABLE I summarizes the performance of our classifier with different types of neural networks and target strength func- tions. The four typical λ functions in FIG 3 are also evalu- ated to demonstrate how these somewhat "pathological" target strength functions affect classification. By Eq. (4), the classification of handwritten digit im- ages could be simplified to a task of restricted linear classification23: given ten classes each with its discriminative (cid:124) i X where X,X∈[0,1]64, image X is classi- function δi(X)= X fied to the class with the largest δi value. Our neural classi- fier simply takes over the computation of vectors' dot product (cid:124) X i X and adds randomness to the ten results. To parameterize the ten δi with their Xi, the "supervisors" could train the neural networks in classifier with the images they deem best -- "av- erage images" in our case or digits learning cards in teachers' case. Our neural classifier is rather unreliable and primitive compared to ANN which is also capable of linear classifica- tion. On one hand, given the same image ANN always outputs the same prediction result. On the other hand, ANN is not only a classifier but also more importantly a "learner", which learns from all kinds of handwritten digits to find the opti- mal Xi for the ten δi; ANN with optimal Xi is more tolerant with poor handwriting, and thus has less misclassification and better prediction accuracy. Only, ANN's learning optimal Xi, an optimization process of many iterations, requires massive computational power to carry out, which is unlikely to be pro- vided by the real-life nervous system -- there is no evidence that an individual neuron can even conduct basic arithmetic operations. Despite of its weakness, our neural classifier has merit in its biological nature: it reduces the computation of vectors' dot product to simple counting of synaptic connec- tions propagated; its training and testing could be purely neu- robiological development and activities where no arithmetic operation is involved; its classification criteria, i.e. "deciding" or "feeling" which neural (sub)network has the most connec- tions propagated, could be an intrinsic capability of intelligent agents. This classifier might project new insights on the neural reality, hopefully. V. CONCLUSION This paper proposes a mathematical theory to explain how memory forms and works. It all begins with synaptic plas- ticity. We find out that, synaptic plasticity is more than im- pulses affecting synapses; it actually plays as a force that can drive neural network eventually to a long-lasting state. We also find out that, under certain conditions there would be a one-to-one mapping between the neural state and the external stimulus that neural network is exposed to. With the map- ping, given stimulus we know exactly what neural state will be; given neural state we know precisely what stimulus has been. The mapping is essentially a link between past event and neural present; between the short-lived and the enduring. In that sense, the mapping itself is memory, or the mapping casts memory in neural network. Next, we study how memory af- fects neural network's response to stimulus. We find out that, the neural network with memory of stimulus can response to similar stimulus more intensely than to the stimulus of less similarity, if response intensity is evaluated by the number of synaptic connections propagated by impulses. That is to say, a neural network with memory is able to classify stimulus. To verify this ability, we experiment with the classifier consist- ing of ten neural networks, and they turn out to have consider- able accuracy in classifying the handwritten digit images. The classifier proves that neurons could collectively provide fully biological computation for classification. Our reasoning takes root in the mathematical treatment of synaptic plasticity as target strength function λ from impulse frequency to synaptic strength. We put hypothetical con- straints on this λ function to ensure that the ideal one-to-one mapping exists. Although these constraints are necessary to keep our theory mathematically sound, they raise concerns. Firstly, they could be overly restrictive. Take continuity con- straint for example. Even the discontinuous function of FIG 3(d), whose nonexistent θ function would map certain stimu- lus to any point within a "fixed interval" instead of a specific fixed point as shown in FIG 6, can be a useable λ in our clas- sifier according to TABLE I. In this case, fixed point per se doesn't have to exist, and mere tendency to seek out for it 8 could serve the purpose. Secondly, as discussed in Section II those λ constraints have yet to be supported by neurobiolog- ical evidences. Above all, the evidence that reveals true λ is vital to clarify the uncertainty. VI. ACKNOWLEDGEMENT We thank the anonymous reviewers for their comments that improved the manuscript. 1J. Hughes. Post-tetanic potentiation. Physiol Rev., 38(1):91 -- 113, 1958. 2D. Hebb. The organization of behavior: a neuropsychological theory. John Wiley & Sons, 1949. 3S. Lowel and W. Singer. Selection of intrinsic horizontal connections in the visual cortex by correlated neuronal activity. Science, 255(5041):209 -- 212, 1992. 4G. Berlucchi and H. Buchtel. Neuronal plasticity: historical roots and evo- lution of meaning. Experimental Brain Research, 192(3):307 -- 19, 2009. 5A. Citri and R. Malenka. Synaptic plasticity: multiple forms, functions, and mechanisms. Neuropsychopharmacology, 33:18 -- 41, 2008. 6S. Martin, P. Grimwood, and R. Morris. Synaptic plasticity and memory: an evaluation of the hypothesis. Annu Rev Neurosci., 23(649-711), 2000. 7E. Takeuchi, A. Duszkiewicz, and R. Morris. The synaptic plasticity and memory hypothesis: encoding, storage and persistence. Philos Trans R Soc Lond B Biol Sci., 369(1633), 2014. 8S. Nabavi, R. Fox, C. Proulx, J. Lin, and R. Tsien R. Malinow. Engineering a memory with ltd and ltp. Nature, 511(348 -- 352), 2014. 9Y. Yang, D. Liu, W. Huang, J. Deng, Y. Sun, Y. Zuo, and M. Poo. Selective synaptic remodeling of amygdalocortical connections associated with fear memory. Nat. Neurosci., 19(1348 -- 1355), 2016. 10P. Trettenbrein. The demise of the synapse as the locus of memory: a looming paradigm shift? Front. Syst. Neurosci., 10, 2016. 11J. Langille and R. Brown. The synaptic theory of memory: a historical survey and reconciliation of recent opposition. Front. Syst. Neurosci., 12, 2018. 12C. Laing and G. Lord. Stochastic methods in neuroscience. Oxford Univer- sity Press, 2010. 13G. Deco, E. Rolls, and R. Romo. Stochastic dynamics as a principle of brain function. Progress in neurobiology, 88(1):1 -- 16, 2009. 14T. Branco and T. Staras. The probability of neurotransmitter release: variability and feedback control at single synapses. Nat Rev Neurosci., 10(5):373 -- 83, 2009. 15C. Baldassi, F. Gerace, H. Kappen, C. Lucibello, L. Saglietti, E. Tartaglione, and R. Zecchina. Role of synaptic stochasticity in training low-precision neural networks. Phys. Rev. Lett., 120(26), 2018. 16J. Schmidhuber. Deep learning in neural networks: an overview. Neural Networks, 61:85 -- 117, 2015. 17I. Istratescu. Fixed point theory: an introduction. Springer, 1981. 18Source code can be found at https://github.com/lansiz/neuron. 19C. Shannon. A mathematical theory of communication. The Bell System Technical Journal, 27(3):379 -- 423, 1948. 20G. Buzsaki. Neural syntax: cell assemblies, synapsembles and readers. Neuron, 63(362 -- 385), 2010. 21C. Butler, Y. Wilson, J. Gunnersen, and M. Murphy. Tracking the fear memory engram: discrete populations of neurons within amygdala, hy- pothalamus and lateral septum are specifically activated by auditory fear conditioning. Learn. Mem., 22(370 -- 384), 2015. 22C. Butler, Y. Wilson, J. Oyrer, T. Karle, S. Petrou, J. Gunnersen, M. Mur- phy, and C. Reid. Neurons specifically activated by fear learning in lateral amygdala display increased synaptic strength. eNeuro, 2018. 23T. Hastie, R. Tibshirani, and J. Friedman. The elements of statistical learn- ing. Springer, 2009. 24The dataset of 1797 handwritten digit images can be obtained with Python code "from sklearn import datasets; datasets.load_digits()".
1608.03619
1
1608
2016-08-11T21:09:40
Modular Segregation of Structural Brain Networks Supports the Development of Executive Function in Youth
[ "q-bio.NC" ]
The human brain is organized into large-scale functional modules that have been shown to evolve in childhood and adolescence. However, it remains unknown whether structural brain networks are similarly refined during development, potentially allowing for improvements in executive function. In a sample of 882 participants (ages 8-22) who underwent diffusion imaging as part of the Philadelphia Neurodevelopmental Cohort, we demonstrate that structural network modules become more segregated with age, with weaker connections between modules and stronger connections within modules. Evolving modular topology facilitated network integration, driven by age-related strengthening of hub edges that were present both within and between modules. Critically, both modular segregation and network integration were associated with enhanced executive performance, and mediated the improvement of executive functioning with age. Together, results delineate a process of structural network maturation that supports executive function in youth.
q-bio.NC
q-bio
Modular Segregation of Structural Brain Networks Supports the Development of Executive Function in Youth Graham L. Baum,a Rastko Ciric,a David R. Roalf,a Richard F. Betzel,c Tyler M. Moore,a Russell T. Shinohara,b Ari E. Kahn,c Megan Quarmley,a Philip A. Cook,d Mark A. Elliott,d Kosha Ruparel,a Raquel E. Gur,a Ruben C. Gur,a Danielle S. Bassett,c,e* & Theodore D. Satterthwaite a* Departments of Psychiatry,a Biostatistics and Epidemiology,b Bioengineering,c Radiology,d and Electrical and Systems Engineeringe of the University of Pennsylvania, Philadelphia PA 19104 USA *=Co-senior authors Address correspondence to: Theodore D. Satterthwaite, M.D., M.A. 10th floor, Gates Building Hospital of the University of Pennsylvania 34th and Spruce Street Philadelphia, PA 19104
 SUMMARY: The human brain is organized into large-scale functional modules that have been shown to evolve in childhood and adolescence. However, it remains unknown whether structural brain networks are similarly refined during development, potentially allowing for improvements in executive function. In a sample of 882 participants (ages 8-22) who underwent diffusion imaging as part of the Philadelphia Neurodevelopmental Cohort, we demonstrate that structural network modules become more segregated with age, with weaker connections between modules and stronger connections within modules. Evolving modular topology facilitated network integration, driven by age-related strengthening of hub edges that were present both within and between modules. Critically, both modular segregation and network integration were associated with enhanced executive performance, and mediated the improvement of executive functioning with age. Together, results delineate a process of structural network maturation that supports executive function in youth. KEYWORDS: development, network, connectome, adolescence, executive, module ! 1 INTRODUCTION Modularity is a fundamental feature of complex systems, including social groups, cyber-physical systems, and diverse biological networks (Newman, 2006). A network module is a group of densely interconnected nodes, which often are the basis for specialized subunits of information processing (Sporns and Betzel, 2016). Functional neuroimaging studies have demonstrated that the human brain has a well-defined modular organization, as reflected in the presence of large-scale functional networks (Biswal et al., 1995; Damoiseaux et al., 2006; Power et al., 2011; Yeo et al., 2011). While the exact number and spatial distribution of functional network modules varies somewhat by analytic approach, a remarkable convergence exists across independent datasets and laboratories (Damoiseaux et al., 2006; Power et al., 2011; Yeo et al., 2011). Commonly described modules include somatomotor (Biswal et al., 1995), visual (Corbetta et al., 1998), default mode (Raichle et al., 2001; Buckner et al., 2008), and fronto-parietal systems (Dosenbach et al., 2007; Vincent et al., 2008). While brain network modules emerge very early in life (Fair et al., 2008; van den Heuvel et al., 2015; Thomason et al., 2013), a growing body of work has shown that these functional modules are refined during youth. During childhood and adolescence, functional modules become more distinct: connectivity within modules increases while connectivity between modules declines (Power et al., 2010; Fair et al., 2007; Fair et al., 2008; Fair et al., 2009; Dosenbach et al., 2010; Gu et al., 2015; Satterthwaite et al., 2013b; Supekar et al., 2009; Anderson et al., 2011). Such development allows for functional specialization, reducing interference among systems (Fornito et al., 2012) and facilitating cognitive performance (Hampson et al., 2010). Modularity is particularly relevant for executive function, which relies on co-activation of executive regions and reciprocal suppression of non-executive regions such as the default mode network (Anticevic et al., 2010; Barber et al., 2013; Persson et al., 2007; Satterthwaite et al., 2013a). Thus, available data suggests that development of network modularity may serve as a substrate for the evolution of executive capability during youth. Despite convergent evidence for the developmental emergence of functional network modularity, there is relatively scant data regarding the maturation of underlying structural brain networks that support this functional architecture (Hermundstad et al., 2014). Prior work demonstrates substantial correspondence between functional and structural measures of brain connectivity (Goñi et al., 2014; Honey et al., 2009; Mišić et al., 2016), although structural connections tend to be a subset of densely connected, polysynaptic functional networks (Betzel et al., 2014; Hermundstad et al., 2013). Structural networks in adults are highly modular (Bassett et al., 2010; Bassett et al., 2011), but it remains unknown if this topology evolves substantially during youth. Correspondence between functional and structural data intuitively suggests that, like functional networks, structural networks should become increasingly segregated during development. However, prior studies using relatively small samples report conflicting results, including declining modularity (Chen et al., 2013), increasing modularity (Chen and Deem, 2015; Huang et al., 2015), or no change with age (Hagmann et al., 2010b; Lim et al., 2015). Larger sample sizes may be necessary for resolving the variability of findings reported in previous studies. ! 2 Beyond this mixed data regarding normative developmental trends, the impact of structural network development on cognitive performance remains poorly described. Cognitive capability improves substantially during youth, with executive function undergoing a protracted phase of development throughout adolescence and young adulthood (Gur et al., 2012; Luna et al., 2004). Describing how structural brain networks evolve to support executive function is necessary to understand the basis for many sources of adolescent morbidity and mortality, which are prominently associated with failures of executive function (Shamosh et al., 2008; Casey et al., 2008; Casey et al., 2011). Finally, such data are a prerequisite for studies of neuropsychiatric disorders, which are increasingly understood as disorders of brain development (Insel, 2010; Rapoport et al., 2012; Casey et al., 2014), are marked by executive dysfunction (Shanmugan et al., 2016), and are linked to the disruption of evolving network topology (Alexander-Bloch et al., 2012; Di Martino et al., 2014; Castellanos and Proal, 2012; Voineskos et al., 2010). Here we sought to define the normative development of structural network modules, and delineate the impact of modular maturation on executive functioning. We tested the hypothesis that modules within structural brain networks become more segregated with age, as seen in functional brain networks. Further, we predicted that segregated structural modules would support enhanced executive functioning. To address these hypotheses, we capitalized upon a large sample of 882 youths who completed diffusion imaging as part of the Philadelphia Neurodevelopmental Cohort (PNC), a community-based study of brain development that includes rich neuroimaging and cognitive data (Satterthwaite et al., 2014a; Calkins et al., 2015; Satterthwaite et al., 2015). As described below, results provide novel evidence that structural brain networks undergo a process of modular segregation analogous to prior accounts of functional network development. Critically, these data reveal that the refinement of structural network modules mediate the development of executive function. RESULTS We investigated the evolution of structural brain networks in a sample of 882 youth aged 8-22 who completed neuroimaging as part of the PNC (Figure 1A). As expected, executive function improved markedly with age (Figure 1B). Structural brain networks were constructed using nodes defined based on a parcellation of each subject's structural image into 234 anatomically defined regions; structural connectivity between these nodes was estimated using deterministic tractography (Figure 2). Each network node was assigned a priori to one of the functional network modules defined by Yeo et al. (2011). Although these module partitions were defined in an independent dataset, using a different imaging modality, the modularity quality of the functional partition imposed on subject-level structural connectivity matrices (QYeo) was highly significant (p<1×10-10). Furthermore, data-driven analysis of structural networks using community detection procedures produced network modules that showed significant similarity to the a priori functional modules (p<1×10-10; Figure S1). Segregation of structural network modules increases with age We first sought to understand whether structural network modules became more segregated with age. To do this, we calculated the average participation coefficient for each subject's network. The participation coefficient quantifies the relative balance of a brain region's ! 3 between-module versus within-module connectivity (Guimerà and Nunes Amaral, 2005). Greater modular segregation is therefore indicated by lower participation coefficient values, with reduced between-module and elevated within-module connectivity. We examined the development of modular segregation using a generalized additive model with penalized splines, which allows for statistically rigorous modeling of both linear and non-linear effects while minimizing over-fitting (Wood, 2004; Wood, 2011). Based on emerging evidence that diffusion-weighted imaging measures are systematically biased by motion artifact (Yendiki et al., 2013), we included in-scanner motion as a covariate in all models in addition to participant sex. To ensure that results reflected changes in network topology, rather than global differences in network connectivity, total network strength was also included as a covariate in all analyses (Li et al., 2012). The participation coefficient declined significantly with age (Figure 3A; p<1×10-10), indicating enhanced modular segregation. Developmental increases in modular segregation were differentially distributed across modules (Figure 3B), with the most robust declines observed in the somatomotor and default mode modules. To further understand which regions were driving these effects, we examined the participation coefficient of individual network nodes. As expected, the nodal participation coefficient declined in many regions (Figure 3C), with many of the most significant reductions occurring in regions within the default mode system. Two exceptions to this overall trend were observed, with increasing participation coefficient in the right rostral frontal gyrus and frontal operculum. Next, we investigated the degree to which developmental effects on modular segregation were driven by changes in within-module connectivity, between-module connectivity, or both. We found that both effects were significant: within-module connectivity increased with age (Figure 4A; p<1×10-10), whereas between-module connectivity declined (Figure 4B; p<1×10-10). Moreover, modular segregation was reflected in individual network edges (Figure 4C), with permutation-based analysis revealing that a higher proportion of connections that strengthened with age were located within a module (Figure 4D; p<0.001). Results are robust to methodological approach Given this strong evidence for developmental modular segregation, we next pursued a set of analyses to determine if our results were dependent upon specific methodological choices. We evaluated multiple parameters, including alternative measures of network segregation, higher-resolution node systems, and different edge measures. First, we examined a complementary measure of modular segregation: the modularity quality index of the Yeo et al. partition (QYeo) applied to individual subject-level structural connectivity matrices. For this metric, greater modular segregation is denoted by a greater QYeo value. As expected, this analysis revealed that modular segregation increased with age (Figure 5A; p=1.06×10-9). Second, while our primary results used the functional partition defined by Yeo et al. (2011), we additionally evaluated modular segregation using a data-driven partition of the structural connectome (see Figures S1 and S2). As seen in Figure 5B, age-related decline in the mean participation coefficient remained evident (p<1×10-10). Third, we directly calculated the modularity quality index of a data-driven partition of each subject's structural connectivity matrix (Qsubj). We observed that this statistic – which provides an individualized measure of modularity that is not dependent on a group-level partition – also increased significantly with age (Figure 5C; p=0.0007), indicating greater modular segregation. Fourth, we investigated the impact of using a more fine-grained ! 4 network parcellation (n=463 nodes). This did not impact the observed results, with age-related declines in the participation coefficient remaining highly significant (Figure 5D; p<1×10-10). Fifth, we evaluated developmental effects when using different measures of structural connectivity instead of mean fractional anisotropy. Results using raw streamline count (Figure 5E; p=6.52×10-7) or volume-normalized streamline density (Figure 5F; p=4.56×10-8) remained highly similar. Lastly, in order to rule out the possibility that observed results were driven by potentially confounding variables, we included additional model covariates such as total brain volume, handedness, race, and maternal education; results were unchanged (p<1×10-10). Conversely, results were consistent when covariates such as total network strength, sex, and motion were removed from the model (p<1×10-10). Modular segregation contributes to global network efficiency Having established that network modules become more segregated with age, and that this finding was not dependent on specific analytic choices, we evaluated the impact of evolving network modularity on measures of network integration. Global network efficiency (Eglob) provides a measure of network integration by quantifying information flow across a network as the shortest path between pairs of nodes (Bassett et al., 2009). In many networks, modularity and global efficiency are inversely related, as network segregation by module partitions extends the path length. However, in some cases it is possible for networks to become both more modular and more efficient; this unusual situation occurs when connectivity within modules is efficiently organized and hub edges form strong links between otherwise segregated modules (Sporns and Betzel, 2016). To determine which scenario characterized human neurodevelopment, we first examined the relationship between global efficiency and age while controlling for the covariates described previously. Replicating previous reports (Chen et al., 2013; Hagmann et al., 2010b), we found that global efficiency increases with age (Figure 6A; p<1×10-10). Next, we calculated the correlation between modular segregation (mean participation coefficient) and global efficiency, while co-varying for age to control for shared developmental trends. Mean participation coefficient was negatively associated with global efficiency (Figure 6B; p<1×10-10), suggesting that the development of network modules does not result in fragmentation, but rather is associated with global network integration. Age effects are concentrated in hub edges that promote network integration To better understand this highly specialized association between network modularity and efficiency, we evaluated the edge betweenness centrality for each network connection. Edge betweenness identifies hub connections by providing a measure of how much a given network edge lies upon the shortest path of communication through a network, and thus contributes to global efficiency. Here we defined hub edges as those connections within the top quartile of edge betweenness across all network edges. Critically, edges that strengthened with age were enriched for hub edges (p<0.001; see Figure 6C). Both within- (p<0.001) and between-module (p<0.001) edges that strengthened with age had higher betweenness than expected by chance (Figure 6D; see Supplemental Information). Furthermore, the average strength of all within-module (Figure 6E; p<1×10-10) and between-module (Figure 6F; p<1×10-10) edges that strengthen with age was associated with global efficiency, suggesting that developmental effects are concentrated within connections that facilitate network integration. The striking combination of increasing modular segregation and enhanced global efficiency demonstrates that structural brain networks ! 5 become more modular and more integrated in development. These dual processes are driven by selective strengthening of network hub edges, which are present within network modules and also provide critical links between increasingly segregated modules. Modular segregation mediates development of executive function in youth Next, we evaluated the cognitive implications of modular segregation by examining associations with individual differences in executive function. Mean whole-brain participation coefficient was associated with improved executive performance (p=0.018). At the level of individual modules, we found that segregation of the frontoparietal control system was uniquely associated with executive ability (Figure 7A; p=0.005), suggesting a network-specific substrate for executive function. As a final step, we examined whether age-related changes in executive function and modularity were related. Mediation analyses revealed that this was indeed the case (Figure 7B; p=0.006), suggesting that the development of segregated structural brain modules mediates the age-related improvement in executive function. These mediating effects were specifically driven by the frontoparietal module (p=0.012). Similarly, global efficiency was associated with executive functioning (p=0.037), and also mediated executive development (p=0.002). To evaluate the specificity of these results, we examined associations with other domains of cognition, such as social cognition and memory performance. While no association with memory was found, modular segregation was also significantly associated with social cognition (p=0.022), which was driven by segregation of the default mode module (p=0.012). This effect mediated improvements in social cognition with age (p=0.008). Together, these results demonstrate that developmental segregation of specific structural network modules may support the development of disparate cognitive domains. DISCUSSION Capitalizing on a large sample of youth imaged as part of the PNC, we demonstrated that modules within human structural brain networks become increasingly segregated with age. This result was robust to specific methodological choices, and driven by a combination of enhanced within-module connectivity and declining between-module connectivity. Age related changes were concentrated within specific hub edges, allowing for networks to simultaneously become more modular and more integrated with age. Critically, segregation of network modules mediated the development of executive function during adolescence. Segregation of structural network modules parallels development of functional networks The delineation of robust, reproducible large-scale functional networks has had a tremendous impact on human neuroscience research (Power et al., 2011; Yeo et al., 2011). As a result, functional network modules have evolved to become the dominant framework by which human imaging data is interpreted. The conceptualization of the brain as a modular entity has had a particularly pronounced effect on theories of development, where convergent results have shown that functional network modules are present early in life (Thomason et al., 2013; van den ! 6 Heuvel et al., 2015), and continue to develop during youth (Power et al., 2010; Fair et al., 2007; Fair et al., 2008; Fair et al., 2009; Dosenbach et al., 2010; Gu et al., 2015; Satterthwaite et al., 2013b; Supekar et al., 2009; Anderson et al., 2011). In contrast, smaller studies of structural brain networks have produced heterogeneous results regarding the development of structural network modules that have not aligned well with functional imaging data (Chen et al., 2013; Chen and Deem, 2015; Hagmann et al., 2010b; Huang et al., 2015; Lim et al., 2015). When considered in light of prior studies that have reported substantial correspondence between brain structure and function (Goñi et al., 2014; Honey et al., 2009; Mišić et al., 2016), the disparity between developmental accounts of structural and functional network modules has been difficult to reconcile. Leveraging a large sample imaged as part of the PNC, we were able to resolve this discrepancy by demonstrating that structural network modules develop in a similar manner as functional brain networks, and become increasingly segregated with age. Modular segregation was present at every scale evaluated, including the whole network, individual network modules, and specific network nodes. Importantly, results were robust to a variety of analytic choices regarding network nodes, edges, and modules; such methodological replication is critical as parameter choices may sometimes impact inference (Hagmann et al., 2010a). For example, while we employed a commonly used set of functional network modules which were defined a priori, analysis of data-driven structural modules provided highly convergent results. Follow-up analyses revealed further parallels with functional imaging studies, and demonstrated that the process of modular segregation is driven by a combination of enhanced connectivity within a module as well as diminished connectivity between modules. Notably, the network modules that demonstrated the greatest developmental segregation were the somatomotor and default mode modules. Prior work has shown that both the default mode and somatomotor systems are highly segregated systems, with a low participation coefficient (Power et al., 2011). The relative isolation of these specific systems in the network may reflect a high degree of processing specialization (Power et al., 2013). The present results thus suggest that the differential evolution of structural network modules is similarly driven by each module's network role. Modular networks become more integrated through strengthening of hub edges In many networks, modular segregation is associated with reduced network integration, as measured by global efficiency. We found that this was not the case in development, and that increasing modularity was in fact associated with enhanced network integration. This robust association was the result of targeted strengthening of specific hub edges. These hub edges were present within but also between modules, allowing for integration across increasingly segregated partitions. These results are congruent with prior studies that have demonstrated that connections between network hubs strengthen preferentially with age (Baker et al., 2015), and that network efficiency increases during development (Chen et al., 2013; Hagmann et al., 2010b). The present data emphasize that increasing modular segregation does not result in isolation of functional sub-systems, but is associated with global network integration through strengthening of hub edges that facilitate both intra- and inter-module connectivity. ! 7 Structural network maturation supports the development of executive function in youth Having defined a normative process of modular segregation, we evaluated the cognitive impact of this developmental effect. While controlling for age, we found that greater modular segregation of structural networks was associated with better executive performance. Critically, modular segregation mediated the observed improvement of executive performance with age, and was driven by segregation of the frontoparietal module. Associations between module segregation and cognition were domain-specific: segregation of the default mode mediated age-related improvements in social cognition, which is reliant on regions within that network (Buckner and Carroll, 2007; Schacter et al., 2007). The process of structural network segregation may allow for functional specialization, and reduce competitive interference between brain systems (Fornito et al., 2012). Such a process is suggested by convergent data from task-based fMRI, which has shown that individual differences in performance are related to selective recruitment of executive regions and suppression of activity elsewhere (Anticevic et al., 2010; Barber et al., 2013; Persson et al., 2007; Satterthwaite et al., 2013a). Furthermore, building on prior work that reported an association between intelligence and the global efficiency of structural (Li et al., 2009) and functional networks (van den Heuvel et al., 2009) in relatively small adult samples, we found that global efficiency also mediated developmental improvements in executive function. Taken together, the current data suggest that structural brain networks re-configure with age, becoming both more modular and more integrated. This specific topology may allow for both functional specialization within modules as well as coordination across modules, which is necessary for effective implementation of dynamic executive processes (Hutchison and Morton, 2015; Braun et al., 2015). Limitations Notwithstanding the strengths of this study, several limitations should be noted. First and foremost, this is a cross-sectional dataset, which has inherent limitations for studies of development (Kraemer et al., 2000). The mediating role that network maturation plays in the development of executive function could be further interrogated using longitudinal data. These limitations offer clear directions for additional research. Ongoing follow-up of this cohort will yield informative data, as will other large-scale studies of brain development, including the IMAGEN consortium (Schumann et al., 2010), the NKI-Rockland sample (Nooner et al., 2012), and the forthcoming Adolescent Brain and Cognitive Development Study. Finally, it should be noted that diffusion-based tractography remains limited in its ability to fully resolve the complex architecture of the structural connectome (Jbabdi et al., 2015). Conclusions In this report, we demonstrated that structural brain modules become more segregated with age. Strengthening of specific within- and between-module hub edges allowed for a simultaneous process of network integration that evolves in concert with modular segregation. Finally, both modular segregation and global network integration mediated the development of executive function in youth. These data resolve an ongoing debate in the field regarding the normative development of structural brain networks, and delineate an important new mechanism for the development of executive functioning in youth. These findings may be relevant for understanding how individual differences in brain development associate with risk-taking behaviors, which are linked to failures of executive function, and are a major source of morbidity ! 8 and mortality in adolescence (Shamosh et al., 2008; Casey et al., 2008; Casey et al., 2011). Furthermore, as both abnormalities within developing networks and executive system dysfunction (Shanmugan et al., 2016) are a common feature of diverse types of psychopathology (Alexander-Bloch et al., 2012; Di Martino et al., 2014; Castellanos and Proal, 2012; Voineskos et al., 2010), structural network development may evolve to become an important imaging biomarker of risk and resilience during the critical period of adolescence. ! 9 EXPERIMENTAL PROCEDURES Participants This study included 882 subjects between 8 and 22 years of age (mean age=15.06, SD=3.15; 389 males, 493 females) who were imaged as part of the PNC (Satterthwaite et al., 2014a; Satterthwaite et al., 2015). Participants were excluded from analyses due to gross structural brain abnormalities (Gur et al., 2013), a history of inpatient psychiatric treatment, current use of psychotropic medications, and medical disorders that could impact brain function. Participants were only included if both structural (Vandekar et al., 2015) and diffusion images (Roalf et al., 2016) passed rigorous quality assurance procedures. All participants completed the Penn computerized neurocognitive battery, which included 14 tests (Gur et al., 2002; Gur et al., 2012). Cognitive performance was summarized by a recent factor analysis (Moore et al., 2014) of both speed and accuracy data, which delineated three factors corresponding to the efficiency of executive function, episodic memory, and social cognition. For further details see Supplemental Information. Image acquisition & processing All high-resolution structural and 64-direction diffusion images were collected on the same 3T Siemens TIM Trio scanner using the same sequences (Satterthwaite et al., 2014a). The T1 image was processed using FreeSurfer version 5.3 (Fischl, 2012), and parcellated into 234 cortical and subcortical regions (Cammoun et al., 2012). The diffusion images were distortion corrected, skull stripped, and motion and eddy current corrected with FSL's `eddy` tool (Jenkinson et al., 2012). Native-space T1 parcels were dilated by 4mm to extend regions into white matter, and co-registered to the first non-weighted (b=0) volume using a boundary-based rigid body transformation (Greve and Fischl, 2009). The diffusion tensor was estimated in DSI Studio and whole-brain deterministic fiber tracking was implemented with 1,000,000 streamlines per subject after removing all streamlines with length less than 10mm (Yeh et al., 2013). Edge weights in the adjacency matrix were defined by mean fractional anisotropy along streamlines connecting each pair of nodes (Mišić et al., 2016; Bohlken et al., 2016; Baker et al., 2015). See Figure 2 and Supplemental Information for more detail. Measurement of modular segregation Nodes were assigned to modules according to their overlap with seven functional networks defined a priori (Yeo et al., 2011); subcortical nodes were assigned to their own additional module. In addition to such a priori functional assignment, network modules were defined directly from the structural connectivity data using a generalized version of the Louvain community detection algorithm (Bassett et al., 2013; Blondel et al., 2008; Mucha et al., 2010). The similarity of data-driven partitions of the structural data were compared to the functional partition using the z-score of the Rand coefficient (Traud et al., 2011). Modular segregation was quantified using the participation coefficient, which measures the balance of between- versus within-module connectivity for each brain region (Guimerà and Nunes Amaral, 2005; Rubinov and Sporns, 2010). A higher participation coefficient at a given node indicates more between-module and less within-module connectivity. Each node's participation coefficient was averaged over modules and the whole brain in order to evaluate modular segregation at multiple scales. To ! 10 understand whether changes in within- or between-module connectivity were driving observed effects, we also calculated these measures separately for the whole brain and each network module (Gu et al., 2015). Also see Supplemental Information. Statistical analyses of age-related changes in modular segregation Linear and nonlinear effects of age were flexibly modeled with penalized thin-plate splines using generalized additive models (GAMs; Wood, 2004; Wood, 2011), which provide statistically rigorous analysis of non-linear effects while minimizing over-fitting. As in previous work (Satterthwaite et al., 2014b; Vandekar et al., 2015), nonlinearity was penalized within the GAMs using restricted maximum likelihood. Developmental models of each network measure were assessed by modeling age with a spline term, while including participant sex and in-scanner motion as covariates. To ensure that changes in global connectivity strength did not drive analyses of network topology, total network strength was also included as a covariate in all analyses (Li et al., 2012). Throughout, multiple comparisons were controlled using the False Discovery Rate (q<0.05; Genovese et al., 2002). Permutation testing was used to assess whether the a priori functional network partition fit subject-level structural connectivity matrices and whether within-module connections were enriched for age effects (see Supplemental Information). Methodological replications To verify that observed age-related increases in modular segregation were not simply due to specific processing choices, we repeated analyses of global network segregation using a variety of other parameters. Analyses were repeated using a subject-specific measure of modularity quality (Q) for both a priori functional networks and a data-driven structural partition. Using the functional network partition, we also examined age-related changes in modular segregation in a higher resolution node system (n=463 instead of n=234), and using two alternative edge definitions (streamline count and normalized streamline density). Finally, we evaluated the effect of additional model covariates (race, maternal education, handedness, and total brain volume). For additional details, see Supplemental Information. Relationship between modular segregation and global network efficiency Global efficiency was calculated for each participant's structural network (Latora and Marchiori, 2001; Rubinov and Sporns, 2010) and age-related effects were examined as above. The relationship between global efficiency and modular segregation was examined while controlling for age (and other covariates as above). To further describe how age-related changes in within- and between-module connectivity might drive measures of global efficiency, we calculated the edge betweenness centrality (Brandes, 2001; Rubinov and Sporns, 2010). Edge betweenness quantifies the degree to which each edge participates in the shortest paths within a network, and thus contributes to global efficiency. Normalized edge betweenness was calculated for each edge, split by type (within- versus between-module) and age effect (strengthens with age versus no change). Hub edges were defined as connections within the top quartile of edge betweenness across all network edges. Permutation tests were used to evaluate whether connections that strengthened with age were enriched for hub edges, and had higher edge betweenness than expected by chance (see Supplemental Information). ! 11 Associations with executive function To determine if network segregation was related to executive performance, a GAM was used which incorporated the factor score for executive efficiency as well as model covariates (spline age, sex, in-scanner motion, total network strength). Specificity analyses additionally evaluated relationships with other cognitive domains including episodic memory and social cognition. Global efficiency was also assessed for relationships with cognition. Linear mediation analyses investigated whether age-related improvements in cognition were mediated by modular segregation or integration; significance was assessed using bootstrapping procedures (Preacher and Hayes, 2008). SUPPLEMENTAL INFORMATION Supplemental Information includes Figure S1, Figure S2, and Supplemental Experimental Procedures. ACKNOWLDGEMENTS We thank the acquisition and recruitment team, including Karthik Prabhakaran and Jeff Valdez. Thanks to Chad Jackson for data management and systems support and Monica Calkins for phenotyping expertise. Supported by grants from the National Institute of Mental Health: R01MH107703 (TDS), R21MH106799 (DSB & TDS). DSB acknowledges support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Research Laboratory and the Army Research Office through contract numbers W911NF-10-2-0022 and W911NF-14-1-0679, the National Institute of Mental Health (R01-DC-009209-11), the National Institute of Child Health and Human Development (R01HD086888-01), the Office of Naval Research, and the National Science Foundation (BCS-1441502, BCS-1430087, and PHY-1554488). The PNC was funded through NIMH RC2 grants MH089983 and MH089924 (REG). Additional support was provided by R01MH107235 (RCG), K23MH098130 (TDS), K01MH102609 (DRR), P50MH096891 (REG), R01NS085211 (RTS), and the Dowshen Program for Neuroscience. Support for developing statistical analyses (RTS & TDS) was provided by a seed grant by the Center for Biomedical Computing and Image Analysis (CBICA) at Penn. Data deposition: The data reported in this paper have been deposited in the database of Genotypes and Phenotypes (dbGaP), www.ncbi.nlm.nih.gov/gap (accession no. phs000607.v1.p1). ! 12 REFERENCES Alexander-Bloch, A., Lambiotte, R., Roberts, B., Giedd, J., Gogtay, N., and Bullmore, E. (2012). The discovery of population differences in network community structure: new methods and applications to brain functional networks in schizophrenia. Neuroimage 59, 3889-3900. Anderson, J.S., Ferguson, M.A., Lopez-Larson, M., and Yurgelun-Todd, D. (2011). Connectivity gradients between the default mode and attention control networks. Brain Connect 1, 147-157. Anticevic, A., Van Snellenberg, J.X., Cohen, R.E., Repovs, G., Dowd, E.C., and Barch, D.M. (2010). Amygdala Recruitment in Schizophrenia in Response to Aversive Emotional Material: A Meta-analysis of Neuroimaging Studies. Schizophr Bull Baker, S.T.E., Lubman, D.I., Yücel, M., Allen, N.B., Whittle, S., Fulcher, B.D., Zalesky, A., and Fornito, A. (2015). Developmental Changes in Brain Network Hub Connectivity in Late Adolescence. J Neurosci 35, 9078-9087. Barber, A.D., Caffo, B.S., Pekar, J.J., and Mostofsky, S.H. (2013). Developmental changes in within- and between-network connectivity between late childhood and adulthood. Neuropsychologia 51, 156-167. Bassett, D.S., Brown, J.A., Deshpande, V., Carlson, J.M., and Grafton, S.T. (2011). Conserved and variable architecture of human white matter connectivity. Neuroimage 54, 1262-1279. Bassett, D.S., Bullmore, E.T., Meyer-Lindenberg, A., Apud, J.A., Weinberger, D.R., and Coppola, R. (2009). Cognitive fitness of cost-efficient brain functional networks. Proc Natl Acad Sci U S A 106, 11747-11752. Bassett, D.S., Greenfield, D.L., Meyer-Lindenberg, A., Weinberger, D.R., Moore, S.W., and Bullmore, E.T. (2010). Efficient physical embedding of topologically complex information processing networks in brains and computer circuits. PLoS Comput Biol 6, e1000748. Bassett, D.S., Porter, M.A., Wymbs, N.F., Grafton, S.T., Carlson, J.M., and Mucha, P.J. (2013). Robust detection of dynamic community structure in networks. Chaos 23, 013142. Betzel, R.F., Byrge, L., He, Y., Goñi, J., Zuo, X.-N., and Sporns, O. (2014). Changes in structural and functional connectivity among resting-state networks across the human lifespan. Neuroimage 102 Pt 2, 345-357. Biswal, B., Yetkin, F.Z., Haughton, V.M., and Hyde, J.S. (1995). Functional connectivity in the motor cortex of resting human brain using echo-planar MRI. Magn Reson Med 34, 537-541. Blondel, V.D., Guillaume, J.-L., Lambiotte, R., and Lefebvre, E. (2008). Fast unfolding of communities in large networks. Journal of statistical mechanics: theory and experiment 2008, P10008. Bohlken, M.M., Brouwer, R.M., Mandl, R.C.W., Van den Heuvel, M.P., Hedman, A.M., De Hert, M., Cahn, W., Kahn, R.S., and Hulshoff Pol, H.E. (2016). Structural Brain Connectivity as a Genetic Marker for Schizophrenia. JAMA Psychiatry 73, 11-19. Braun, U., Schäfer, A., Walter, H., Erk, S., Romanczuk-Seiferth, N., Haddad, L., Schweiger, J.I., Grimm, O., Heinz, A., and Tost, H. (2015). Dynamic reconfiguration of frontal brain ! 13 networks during executive cognition in humans. Proceedings of the National Academy of Sciences 112, 11678-11683. Buckner, R.L., and Carroll, D.C. (2007). Self-projection and the brain. Trends Cogn Sci 11, 49-57. Buckner, R.L., Andrews-Hanna, J.R., and Schacter, D.L. (2008). The brain's default network: anatomy, function, and relevance to disease. Ann N Y Acad Sci 1124, 1-38. Calkins, M.E., Merikangas, K.R., Moore, T.M., Burstein, M., Behr, M.A., Satterthwaite, T.D., Ruparel, K., Wolf, D.H., Roalf, D.R., Mentch, F.D., Qiu, H., Chiavacci, R., Connolly, J.J., Sleiman, P.M.A., Gur, R.C., Hakonarson, H., and Gur, R.E. (2015). The Philadelphia Neurodevelopmental Cohort: constructing a deep phenotyping collaborative. J Child Psychol Psychiatry Cammoun, L., Gigandet, X., Meskaldji, D., Thiran, J.P., Sporns, O., Do, K.Q., Maeder, P., Meuli, R., and Hagmann, P. (2012). Mapping the human connectome at multiple scales with diffusion spectrum MRI. J Neurosci Methods 203, 386-397. Casey, B.J., Hare, T.A., and Galván, A. (2011). Risky and impulsive components of adolescent decision making. Decision Making, Affect, and Learning: Attention and Performance XXIII 23, 425. Casey, B.J., Jones, R.M., and Hare, T.A. (2008). The Adolescent Brain. Ann N Y Acad Sci 1124, 111-126. Casey, B.J., Oliveri, M.E., and Insel, T. (2014). A Neurodevelopmental Perspective on the Research Domain Criteria (RDoC) Framework. Biol Psychiatry 76, 350-353. Castellanos, F.X., and Proal, E. (2012). Large-scale brain systems in ADHD: beyond the prefrontal–striatal model. Trends Cogn Sci 16, 17-26. Chen, M., and Deem, M.W. (2015). Development of modularity in the neural activity of children's brains. Phys Biol 12, 016009. Chen, Z., Liu, M., Gross, D.W., and Beaulieu, C. (2013). Graph theoretical analysis of developmental patterns of the white matter network. Front Hum Neurosci 7, 716. Corbetta, M., Akbudak, E., Conturo, T.E., Snyder, A.Z., Ollinger, J.M., Drury, H.A., Linenweber, M.R., Petersen, S.E., Raichle, M.E., and Van Essen, D.C. (1998). A common network of functional areas for attention and eye movements. Neuron 21, 761-773. Damoiseaux, J.S., Rombouts, S.A.R.B., Barkhof, F., Scheltens, P., Stam, C.J., Smith, S.M., and Beckmann, C.F. (2006). Consistent resting-state networks across healthy subjects. Proc Natl Acad Sci U S A 103, 13848-13853. Di Martino, A., Fair, D.A., Kelly, C., Satterthwaite, T.D., Castellanos, F.X., Thomason, M.E., Craddock, R.C., Luna, B., Leventhal, B.L., Zuo, X.N., and Milham, M.P. (2014). Unraveling the Miswired Connectome: A Developmental Perspective. Neuron 83, 1335-1353. Dosenbach, N.U., Fair, D.A., Miezin, F.M., Cohen, A.L., Wenger, K.K., Dosenbach, R.A., Fox, M.D., Snyder, A.Z., Vincent, J.L., Raichle, M.E., Schlaggar, B.L., and Petersen, S.E. (2007). Distinct brain networks for adaptive and stable task control in humans. Proc Natl Acad Sci U S A 104, 11073-11078. Dosenbach, N.U.F., Nardos, B., Cohen, A.L., Fair, D.A., Power, J.D., Church, J.A., Nelson, S.M., Wig, G.S., Vogel, A.C., Lessov-Schlaggar, C.N., Barnes, K.A., Dubis, J.W., Feczko, E., Coalson, R.S., Pruett, J.R., Barch, D.M., Petersen, S.E., and Schlaggar, B.L. (2010). Prediction of individual brain maturity using fMRI. Science 329, 1358-1361. ! 14 Fair, D.A., Cohen, A.L., Dosenbach, N.U.F., Church, J.A., Miezin, F.M., Barch, D.M., Raichle, M.E., Petersen, S.E., and Schlaggar, B.L. (2008). The maturing architecture of the brain's default network. Proceedings of the National Academy of Sciences 105, 4028. Fair, D.A., Cohen, A.L., Power, J.D., Dosenbach, N.U.F., Church, J.A., Miezin, F.M., Schlaggar, B.L., and Petersen, S.E. (2009). Functional brain networks develop from a "local to distributed" organization. PLoS Comput Biol 5, e1000381. Fair, D.A., Dosenbach, N.U.F., Church, J.A., Cohen, A.L., Brahmbhatt, S., Miezin, F.M., Barch, D.M., Raichle, M.E., Petersen, S.E., and Schlaggar, B.L. (2007). Development of distinct control networks through segregation and integration. Proc Natl Acad Sci U S A 104, 13507-13512. Fischl, B. (2012). FreeSurfer. Neuroimage 62, 774-781. Fornito, A., Harrison, B.J., Zalesky, A., and Simons, J.S. (2012). Competitive and cooperative dynamics of large-scale brain functional networks supporting recollection. Proc Natl Acad Sci U S A 109, 12788-12793. Genovese, C.R., Lazar, N.A., and Nichols, T. (2002). Thresholding of statistical maps in functional neuroimaging using the false discovery rate. Neuroimage 15, 870-878. Goñi, J., van den Heuvel, M.P., Avena-Koenigsberger, A., de Mendizabal, N.V., Betzel, R.F., Griffa, A., Hagmann, P., Corominas-Murtra, B., Thiran, J.-P., and Sporns, O. (2014). Resting-brain functional connectivity predicted by analytic measures of network communication. Proceedings of the National Academy of Sciences 111, 833-838. Greve, D.N., and Fischl, B. (2009). Accurate and robust brain image alignment using boundary-based registration. Neuroimage 48, 63-72. Gu, S., Satterthwaite, T.D., Medaglia, J.D., Yang, M., Gur, R.E., Gur, R.C., and Bassett, D.S. (2015). Emergence of system roles in normative neurodevelopment. Proc Natl Acad Sci U S A 112, 13681-13686. Guimerà, R., and Nunes Amaral, L.A. (2005). Functional cartography of complex metabolic networks. Nature 433, 895-900. Gur, R.C., Richard, J., Calkins, M.E., Chiavacci, R., Hansen, J.A., Bilker, W.B., Loughead, J., Connolly, J.J., Qiu, H., Mentch, F.D., Abou-Sleiman, P.M., Hakonarson, H., and Gur, R.E. (2012). Age group and sex differences in performance on a computerized neurocognitive battery in children age 8-21. Neuropsychology 26, 251-265. Gur, R.C., Sara, R., Hagendoorn, M., Marom, O., Hughett, P., Macy, L., Turner, T., Bajcsy, R., Posner, A., and Gur, R.E. (2002). A method for obtaining 3-dimensional facial expressions and its standardization for use in neurocognitive studies. J Neurosci Methods 115, 137-143. Gur, R.E., Kaltman, D., Melhem, E.R., Ruparel, K., Prabhakaran, K., Riley, M., Yodh, E., Hakonarson, H., Satterthwaite, T., and Gur, R.C. (2013). Incidental findings in youths volunteering for brain MRI research. AJNR Am J Neuroradiol 34, 2021-2025. Hagmann, P., Cammoun, L., Gigandet, X., Gerhard, S., Ellen Grant, P., Wedeen, V., Meuli, R., Thiran, J.-P., Honey, C.J., and Sporns, O. (2010a). MR connectomics: Principles and challenges. J Neurosci Methods Hagmann, P., Sporns, O., Madan, N., Cammoun, L., Pienaar, R., Wedeen, V.J., Meuli, R., Thiran, J.-P., and Grant, P.E. (2010b). White matter maturation reshapes structural connectivity in the late developing human brain. Proc Natl Acad Sci U S A 107, 19067-19072. ! 15 Hampson, M., Driesen, N., Roth, J.K., Gore, J.C., and Constable, R.T. (2010). Functional connectivity between task-positive and task-negative brain areas and its relation to working memory performance. Magn Reson Imaging Hermundstad, A.M., Bassett, D.S., Brown, K.S., Aminoff, E.M., Clewett, D., Freeman, S., Frithsen, A., Johnson, A., Tipper, C.M., and Miller, M.B. (2013). Structural foundations of resting-state and task-based functional connectivity in the human brain. Proceedings of the National Academy of Sciences 110, 6169-6174. Hermundstad, A.M., Brown, K.S., Bassett, D.S., Aminoff, E.M., Frithsen, A., Johnson, A., Tipper, C.M., Miller, M.B., Grafton, S.T., and Carlson, J.M. (2014). Structurally-constrained relationships between cognitive states in the human brain. PLoS Comput Biol 10, e1003591. Honey, C.J., Sporns, O., Cammoun, L., Gigandet, X., Thiran, J.P., Meuli, R., and Hagmann, P. (2009). Predicting human resting-state functional connectivity from structural connectivity. Proc Natl Acad Sci U S A 106, 2035-2040. Huang, H., Shu, N., Mishra, V., Jeon, T., Chalak, L., Wang, Z.J., Rollins, N., Gong, G., Cheng, H., Peng, Y., Dong, Q., and He, Y. (2015). Development of human brain structural networks through infancy and childhood. Cereb Cortex 25, 1389-1404. Hutchison, R.M., and Morton, J.B. (2015). It's a matter of time: Reframing the development of cognitive control as a modification of the brain's temporal dynamics. Dev Cogn Neurosci Insel, T.R. (2010). Rethinking schizophrenia. Nature 468, 187-193. Jenkinson, M., Beckmann, C.F., Behrens, T.E.J., Woolrich, M.W., and Smith, S.M. (2012). FSL. Neuroimage 62, 782-790. Jbabdi, S., Sotiropoulos, S.N., Haber, S.N., Van Essen, D.C., and Behrens, T.E. (2015). Measuring macroscopic brain connections in vivo. Nat Neurosci 18, 1546-1555. Kraemer, H.C., Yesavage, J.A., Taylor, J.L., and Kupfer, D. (2000). How can we learn about developmental processes from cross-sectional studies, or can we? Am J Psychiatry 157, 163-171. Latora, V., and Marchiori, M. (2001). Efficient behavior of small-world networks. Phys Rev Lett 87, 198701. Li, L., Rilling, J.K., Preuss, T.M., Glasser, M.F., and Hu, X. (2012). The effects of connection reconstruction method on the interregional connectivity of brain networks via diffusion tractography. Hum Brain Mapp 33, 1894-1913. Li, Y., Liu, Y., Li, J., Qin, W., Li, K., Yu, C., and Jiang, T. (2009). Brain anatomical network and intelligence. PLoS Comput Biol 5, e1000395. Lim, S., Han, C.E., Uhlhaas, P.J., and Kaiser, M. (2015). Preferential detachment during human brain development: age- and sex-specific structural connectivity in diffusion tensor imaging (DTI) data. Cereb Cortex 25, 1477-1489. Luna, B., Garver, K.E., Urban, T.A., Lazar, N.A., and Sweeney, J.A. (2004). Maturation of cognitive processes from late childhood to adulthood. Child Dev 75, 1357-1372. Mišić, B., Betzel, R.F., de Reus, M.A., van den Heuvel, M.P., Berman, M.G., McIntosh, A.R., and Sporns, O. (2016). Network-Level Structure-Function Relationships in Human Neocortex. Cereb Cortex 26, 3285-3296. Moore, T.M., Reise, S.P., Gur, R.E., Hakonarson, H., and Gur, R.C. (2014). Psychometric Properties of the Penn Computerized Neurocognitive Battery. Neuropsychology ! 16 Mucha, P.J., Richardson, T., Macon, K., Porter, M.A., and Onnela, J.-P. (2010). Community structure in time-dependent, multiscale, and multiplex networks. Science 328, 876-878. Newman, M.E.J. (2006). Modularity and community structure in networks. Proc Natl Acad Sci U S A 103, 8577-8582. Nooner, K.B., Colcombe, S.J., Tobe, R.H., Mennes, M., Benedict, M.M., Moreno, A.L., Panek, L.J., Brown, S., Zavitz, S.T., Li, Q., Sikka, S., Gutman, D., Bangaru, S., Schlachter, R.T., Kamiel, S.M., Anwar, A.R., Hinz, C.M., Kaplan, M.S., Rachlin, A.B., Adelsberg, S., Cheung, B., Khanuja, R., Yan, C., Craddock, C.C., Calhoun, V., Courtney, W., King, M., Wood, D., Cox, C.L., Kelly, A.M.C., Di Martino, A., Petkova, E., Reiss, P.T., Duan, N., Thomsen, D., Biswal, B., Coffey, B., Hoptman, M.J., Javitt, D.C., Pomara, N., Sidtis, J.J., Koplewicz, H.S., Castellanos, F.X., Leventhal, B.L., and Milham, M.P. (2012). The NKI-Rockland Sample: A Model for Accelerating the Pace of Discovery Science in Psychiatry. Front Neurosci 6, 152. Persson, J., Lustig, C., Nelson, J.K., and Reuter-Lorenz, P.A. (2007). Age differences in deactivation: a link to cognitive control? J Cogn Neurosci 19, 1021-1032. Power, J.D., Cohen, A.L., Nelson, S.M., Wig, G.S., Barnes, K.A., Church, J.A., Vogel, A.C., Laumann, T.O., Miezin, F.M., Schlaggar, B.L., and Petersen, S.E. (2011). Functional network organization of the human brain. Neuron 72, 665-678. Power, J.D., Fair, D.A., Schlaggar, B.L., and Petersen, S.E. (2010). The development of human functional brain networks. Neuron 67, 735-748. Power, J.D., Schlaggar, B.L., Lessov-Schlaggar, C.N., and Petersen, S.E. (2013). Evidence for hubs in human functional brain networks. Neuron 79, 798-813. Preacher, K.J., and Hayes, A.F. (2008). Asymptotic and resampling strategies for assessing and comparing indirect effects in multiple mediator models. Behav Res Methods 40, 879-891. Raichle, M.E., MacLeod, A.M., Snyder, A.Z., Powers, W.J., Gusnard, D.A., and Shulman, G.L. (2001). A default mode of brain function. Proc Natl Acad Sci U S A 98, 676-682. Rapoport, J.L., Giedd, J.N., and Gogtay, N. (2012). Neurodevelopmental model of schizophrenia: update 2012. Mol Psychiatry Roalf, D.R., Quarmley, M., Elliott, M.A., Satterthwaite, T.D., Vandekar, S.N., Ruparel, K., Gennatas, E.D., Calkins, M.E., Moore, T.M., Hopson, R., Prabhakaran, K., Jackson, C.T., Verma, R., Hakonarson, H., Gur, R.C., and Gur, R.E. (2016). The impact of quality assurance assessment on diffusion tensor imaging outcomes in a large-scale population-based cohort. Neuroimage 125, 903-919. Rubinov, M., and Sporns, O. (2010). Complex network measures of brain connectivity: uses and interpretations. Neuroimage 52, 1059-1069. Satterthwaite, T.D., Connolly, J.J., Ruparel, K., Calkins, M.E., Jackson, C., Elliott, M.A., Roalf, D.R., Ryan Hopsona, K.P., Behr, M., Qiu, H., Mentch, F.D., Chiavacci, R., Sleiman, P.M.A., Gur, R.C., Hakonarson, H., and Gur, R.E. (2015). The Philadelphia Neurodevelopmental Cohort: A publicly available resource for the study of normal and abnormal brain development in youth. Neuroimage Satterthwaite, T.D., Elliott, M.A., Ruparel, K., Loughead, J., Prabhakaran, K., Calkins, M.E., Hopson, R., Jackson, C., Keefe, J., Riley, M., Mentch, F.D., Sleiman, P., Verma, R., Davatzikos, C., Hakonarson, H., Gur, R.C., and Gur, R.E. (2014a). Neuroimaging of the Philadelphia neurodevelopmental cohort. Neuroimage 86, 544-553. ! 17 Satterthwaite, T.D., Shinohara, R.T., Wolf, D.H., Hopson, R.D., Elliott, M.A., Vandekar, S.N., Ruparel, K., Calkins, M.E., Roalf, D.R., Gennatas, E.D., Jackson, C., Erus, G., Prabhakaran, K., Davatzikos, C., Detre, J.A., Hakonarson, H., Gur, R.C., and Gur, R.E. (2014b). Impact of puberty on the evolution of cerebral perfusion during adolescence. Proc Natl Acad Sci U S A Satterthwaite, T.D., Wolf, D.H., Erus, G., Ruparel, K., Elliott, M.A., Gennatas, E.D., Hopson, R., Jackson, C., Prabhakaran, K., Bilker, W.B., Calkins, M.E., Loughead, J., Smith, A., Roalf, D.R., Hakonarson, H., Verma, R., Davatzikos, C., Gur, R.C., and Gur, R.E. (2013a). Functional Maturation of the Executive System during Adolescence. J Neurosci 33, 16249-16261. Satterthwaite, T.D., Wolf, D.H., Ruparel, K., Erus, G., Elliott, M.A., Eickhoff, S.B., Gennatas, E.D., Jackson, C., Prabhakaran, K., Smith, A., Hakonarson, H., Verma, R., Davatzikos, C., Gur, R.E., and Gur, R.C. (2013b). Heterogeneous impact of motion on fundamental patterns of developmental changes in functional connectivity during youth. Neuroimage 83, 45 - 57. Schacter, D.L., Addis, D.R., and Buckner, R.L. (2007). Remembering the past to imagine the future: the prospective brain. Nat Rev Neurosci 8, 657-661. Schumann, G., Loth, E., Banaschewski, T., Barbot, A., Barker, G., Büchel, C., Conrod, P.J., Dalley, J.W., Flor, H., Gallinat, J., Garavan, H., Heinz, A., Itterman, B., Lathrop, M., Mallik, C., Mann, K., Martinot, J.-L., Paus, T., Poline, J.-B., Robbins, T.W., Rietschel, M., Reed, L., Smolka, M., Spanagel, R., Speiser, C., Stephens, D.N., Ströhle, A., Struve, M., and IMAGEN consortium (2010). The IMAGEN study: reinforcement-related behaviour in normal brain function and psychopathology. Mol Psychiatry 15, 1128-1139. Shamosh, N.A., Deyoung, C.G., Green, A.E., Reis, D.L., Johnson, M.R., Conway, A.R.A., Engle, R.W., Braver, T.S., and Gray, J.R. (2008). Individual differences in delay discounting: relation to intelligence, working memory, and anterior prefrontal cortex. Psychol Sci 19, 904-911. Shanmugan, S., Wolf, D.H., Calkins, M.E., Moore, T.M., Ruparel, K., Hopson, R.D., Vandekar, S.N., Roalf, D.R., Elliott, M.A., Jackson, C., Gennatas, E.D., Leibenluft, E., Pine, D.S., Shinohara, R.T., Hakonarson, H., Gur, R.C., Gur, R.E., and Satterthwaite, T.D. (2016). Common and Dissociable Mechanisms of Executive System Dysfunction Across Psychiatric Disorders in Youth. Am J Psychiatry , appiajp201515060725. Sporns, O., and Betzel, R.F. (2016). Modular Brain Networks. Annu Rev Psychol 67, 613-640. Supekar, K., Musen, M., and Menon, V. (2009). Development of large-scale functional brain networks in children. PLoS Biol 7, e1000157. Thomason, M.E., Dassanayake, M.T., Shen, S., Katkuri, Y., Alexis, M., Anderson, A.L., Yeo, L., Mody, S., Hernandez-Andrade, E., Hassan, S.S., Studholme, C., Jeong, J.-W., and Romero, R. (2013). Cross-hemispheric functional connectivity in the human fetal brain. Sci Transl Med 5, 173ra24. Traud, A.L., Kelsic, E.D., Mucha, P.J., and Porter, M.A. (2011). Comparing community structure to characteristics in online collegiate social networks. SIAM review 53, 526-543. Vandekar, S.N., Shinohara, R.T., Raznahan, A., Roalf, D.R., Ross, M., DeLeo, N., Ruparel, K., Verma, R., Wolf, D.H., Gur, R.C., Gur, R.E., and Satterthwaite, T.D. (2015). Topologically dissociable patterns of development of the human cerebral cortex. J Neurosci 35, 599-609. ! 18 van den Heuvel, M.P., Kersbergen, K.J., de Reus, M.A., Keunen, K., Kahn, R.S., Groenendaal, F., de Vries, L.S., and Benders, M.J.N.L. (2015). The Neonatal Connectome During Preterm Brain Development. Cereb Cortex 25, 3000-3013. van den Heuvel, M.P., Stam, C.J., Kahn, R.S., and Hulshoff Pol, H.E. (2009). Efficiency of functional brain networks and intellectual performance. J Neurosci 29, 7619-7624. Vincent, J.L., Kahn, I., Snyder, A.Z., Raichle, M.E., and Buckner, R.L. (2008). Evidence for a frontoparietal control system revealed by intrinsic functional connectivity. J Neurophysiol 100, 3328-3342. Voineskos, A.N., Lobaugh, N.J., Bouix, S., Rajji, T.K., Miranda, D., Kennedy, J.L., Mulsant, B.H., Pollock, B.G., and Shenton, M.E. (2010). Diffusion tensor tractography findings in schizophrenia across the adult lifespan. Brain 133, 1494-1504. Wood, S.N. (2004). Stable and efficient multiple smoothing parameter estimation for generalized additive models. Journal of the American Statistical Association 99 Wood, S.N. (2011). Fast stable restricted maximum likelihood and marginal likelihood estimation of semiparametric generalized linear models. Journal of the Royal Statistical Society: Series B (Statistical Methodology) 73, 3-36. Yeh, F.-C., Verstynen, T.D., Wang, Y., Fernández-Miranda, J.C., and Tseng, W.-Y.I. (2013). Deterministic diffusion fiber tracking improved by quantitative anisotropy. PLoS One 8, e80713. Yendiki, A., Koldewyn, K., Kakunoori, S., Kanwisher, N., and Fischl, B. (2013). Spurious group differences due to head motion in a diffusion MRI study. Neuroimage 88C, 79-90. Yeo, B.T.T., Krienen, F.M., Sepulcre, J., Sabuncu, M.R., Lashkari, D., Hollinshead, M., Roffman, J.L., Smoller, J.W., Zöllei, L., Polimeni, J.R., Fischl, B., Liu, H., and Buckner, R.L. (2011). The organization of the human cerebral cortex estimated by intrinsic functional connectivity. J Neurophysiol 106, 1125-1165. ! 19 Figure 1. Executive functioning improves with age. A. Age distribution of 882 youth completing diffusion imaging as part of the PNC. B. Executive performance on a neurocognitive battery improves with age. Blue line represents the best fit from a general additive model; shaded area indicates 95% confidence interval. Model includes participant sex as a covariate. ! 20 Figure 2. Connectome construction. For each subject, the T1 image was processed using FreeSurfer and parcellated into 234 network nodes on an individualized basis. Deterministic whole-brain fiber tracking was used to create a symmetric adjacency matrix (234×234), where the edge weight was defined as the mean fractional anisotropy (FA) along the connecting streamlines. Network nodes were each assigned to one of the seven large-scale functional modules defined by Yeo et al. (2011); subcortical nodes were assigned to an eighth modules. VIS=visual, SOM=somatomotor, DOR=dorsal attention, VEN=ventral attention, LIM=limbic, FPC=frontoparietal control, DMN= default mode network, SUB=subcortical. ! 21 Figure 3. Structural brain network modules become increasingly segregated with age. A. Modular segregation was quantified as the mean participation coefficient across all network nodes, with lower values indicating more segregation. Participation coefficient values declined significantly with age. B. Modular segregation is differentially distributed across functional systems. Age-related modular segregation is most robust in the somatomotor and default mode systems, but also present in other networks. C. Age-related changes in participation coefficient provide convergent results for individual nodes, and demonstrate widespread declines with age. The strongest age-related reductions of the participation coefficient were seen in default mode regions such as the posterior cingulate. Two exceptions to this overall trend were the right rostral frontal gyrus and frontal operculum, where participation coefficient increased with age. Blue line represents the best fit from a general additive model; shaded area indicates 95% confidence interval. All analyses control for sex, in-scanner motion, and network strength. Color palette represents z-transformed p values from a general additive model. Images are thresholded to control for multiple comparisons using the False Discovery Rate (Q<0.05). *indicates p<0.001. ! 22 Figure 4. Modular segregation is driven by a combination of both enhanced within-module connectivity and reduced between-module connectivity. A. Average strength of within-module connectivity increases with age. B. Between-module connectivity decreases across development. C. Convergent effects are seen at the level of individual graph edges (image thresholded using Bonferroni corrected p<0.05 for clarity). D. A higher percentage of within-module connections (red) strengthen with age than expected by chance. All analyses include sex, in-scanner motion, and network strength as model covariates. * indicates p<0.001. ! 23 Figure 5. Results are robust to methodological choices. Regardless of specific processing decisions, an increase in modular segregation with age was observed. A. Convergent findings result when using an index of the modularity quality for the Yeo partition, where higher Q indicates more segregated modules. B. When using a group-level structural partition, modular segregation (mean participation coefficient) decreases with age. C. Modularity quality of subject-level connectivity matrices also increases with age. D. Results remain unaffected when a higher-dimensional parcellation is used (n=463 nodes), E. when streamline count is used instead of FA as an edge weight, and F. when normalized streamline density is used as the edge weight. Lower participation coefficient indicates more segregated modules. Blue line represents the best fit from a general additive model; shaded area indicates 95% confidence interval. All models include sex, in-scanner motion, and total network strength as covariates. ! 24 Figure 6. Modularity is associated with network integration, and is driven by developmental strengthening of specific hub edges. A. Replicating prior work, global network efficiency increases with age. Model includes sex, in-scanner motion, and total network strength as covariates. B. While controlling for age, lower mean participation coefficient is associated with greater network efficiency, indicating a positive association between modular segregation and network integration. C. Connections that strengthen with age are enriched for hub edges (47%). Hub edges are defined as connections in the top quartile of edge betweenness centrality, which quantifies how often a given edge lies on the shortest path between nodes and thus facilitates global efficiency. Image thresholded using Bonferroni corrected p<0.05 for clarity. D. Both within- and between-module connections that strengthen with age have higher edge betweenness centrality than expected by chance. The average weight of within- (E) and between-module edges (F) that strengthen with age are positively associated with global efficiency. Blue line represents the best fit from a general additive model, shaded area indicates 95% confidence interval; * indicates p<0.001. Error bar represents standard error of the mean. ! 25 Figure 7. Segregation of structural modules supports the development of executive function in youth. A. While controlling for age, greater modular segregation in the frontoparietal control network is associated with better executive performance. B. Segregation of structural modules mediates the improvement of executive function with age. Mediation results shown as unstandardized regression coefficients. Significance of indirect effect (c'=0.007) was assessed using bootstrapped confidence intervals [0.002-0.012]. All models also include sex, in-scanner motion, and total network strength as covariates. * indicates p<0.01. ! 26 Figure 8. Modular evolution of structural brain networks across youth. From childhood through adulthood, structural brain networks become increasingly modular. The targeted strengthening of specific hub edges facilitates specialized information processing within distinct modules, and simultaneously enhances integration across modules. Hub edges are indicated by thick connections. ! 27 SUPPLEMENTAL INFORMATION! 28 Supplemental Figures Supplemental Figure 1: Number of modules identified in group-level structural par- titions. To examine alternative data-driven modular partitions of structural brain networks, we varied γ over the interval [0,4] in increments of 0.05. The number of modules identified in group-level consensus partitions increases as a function of γ. The similarity between structural partitions and a priori functional partitions also increases with γ and the number of identified structural modules. ∗ indicates alternative structural partitions identified at plateaus for the number of modules. Bars are colored by the z-score of the Rand coeffi- cient, which quantifies the similarity between structural partitions and the a priori func- tional partition used throughout the main text. The 9-module structural partition identified at γ=2.5 (marked by blue box) is used to examine age-related effects on modular segre- gation in Figure 5. The z-score of the Rand coefficient is equal to 17.6 (p < 1× 10−10) for this structural partition, suggesting a significant similarity with the functional partition beyond chance. 30 051015201234GammaRastko_numComms_n882102030Zrand_n882_consensus_vs_Yeo7systemNumber of Modules123405101520102030ZRand*** = 1.5 = 3.1 = 2.5 Supplemental Figure 2: Data-driven structural network modules become more seg- regated across youth. Here we demonstrate that regardless of the group-level consensus partition used to define modules, modular segregation increases with age, as demonstrated by a significant decrease in the mean participation coefficient. This developmental pattern is replicated using a 5-module partition (A, γ=1.5), a 9-module partition (B, γ=2.5), and an 11-module partition (C, γ=3.1). The 9-module partition pictured in B is used to calculate modular segregation in Figure 5. Blue line represents the best fit from a general additive model; shaded area indicates 95% confidence interval. Models include participant sex, in-scanner head motion, and total network strength as covariates. 31 8101214161820220.460.480.500.520.540.560.58age_in_yrsf(age_in_yrs)0.460.480.500.520.540.560.58810121416182022p < 1 × 10-10Age (years)Mean Participation CoefficientModuleModuleModule8101214161820220.670.680.690.700.710.720.730.74age_in_yrsf(age_in_yrs)810121416182022Age (years)0.670.680.690.700.710.720.730.74p < 1 × 10-10Mean Participation Coefficient8101214161820220.600.620.640.660.68age_in_yrsf(age_in_yrs)0.60Mean Participation Coefficient0.620.640.660.68810121416182022Age (years)p < 1 × 10-10 = 1.5 = 2.5 = 3.1234511234567891234567891011 Supplemental Experimental Procedures Subjects Diffusion tensor imaging (DTI) datasets were acquired as part of the Philadelphia Neu- rodevelopmental Cohort (PNC), a large community-based study of brain development. 1601 subjects completed the cross-sectional neuroimaging protocol (Satterthwaite et al., 2014). Datasets from 244 individuals were considered unusable due to incomplete ac- quisition or incidental findings. The remaining 1357 participants underwent a rigorous manual and automated quality assurance protocol for DTI datasets (Roalf et al., 2016), which flagged 157 subjects for poor data quality (e.g., low temporal signal-to-noise ratio). Of the remaining 1210 participants, 93 were flagged by automated quality assurance for low quality or incomplete FreeSurfer reconstruction of T1-weighted images. Of the re- maining 1117 participants, 235 subjects were excluded for meeting any of the following criteria: gross radiological abnormalities, history of medical problems that might affect brain function, history of inpatient psychiatric hospitalization, use of psychotropic medi- cation at the time of data acquisition, missing data, and/or high levels of in-scanner head motion (mean relative displacement between non-weighted volumes > 2mm), which has been shown to impact measures derived from diffusion-weighted imaging (Roalf et al., 2016; Yendiki et al., 2013). These exclusions produced a final sample consisting of 882 youths (mean age=15.06, SD=3.15; 389 males, 493 females). Cognitive Assessment The Penn computerized neurocognitive battery (Penn CNB) was administered to all participants. The CNB consists of 14 tests adapted from tasks applied in functional neu- roimaging to evaluate a broad range of cognitive domains (Gur et al., 2002; Gur et al., 2012). These domains include executive control (abstraction and flexibility, attention, working memory), episodic memory (verbal, facial, spatial), complex cognition (verbal reasoning, nonverbal reasoning, spatial processing), social cognition (emotion identifica- tion, emotion intensity differentiation, age differentiation) and sensorimotor and motor speed. Accuracy and speed for each test were z-transformed. Cognitive performance was summarized by a recent factor analysis (Moore et al., 2014) of both speed and accuracy data, which delineated three factors corresponding to the efficiency of executive function, episodic memory, and social cognition. Data Acquisition All MRI scans were acquired on the same 3T Siemens Tim Trio whole-body scanner and 32-channel head coil at the Hospital of the University of Pennsylvania. DTI scans were acquired using a twice- refocused spin-echo (TRSE) single-shot echo-planar imaging 32 (EPI) sequence (TR = 8100ms, TE = 82ms, FOV = 240mm2 /240mm2 ; Matrix = RL: 128/AP:128/Slices:70, in-plane resolution (x and y) 1.875 mm2; slice thickness = 2mm, gap = 0; flip angle = 90◦/180◦/180◦, volumes = 71, GRAPPA factor = 3, bandwidth = 2170 Hz/pixel, PE direction = AP). This sequence used a four-lobed diffusion encoding gradient scheme combined with a 90-180-180 spin-echo sequence designed to minimize eddy-current artifacts . DTI data were acquired in two consecutive series consisting of 32 diffusion encoding gradient schemes. The complete sequence consisted of 64 diffusion- weighted directions with b=1000s/mm2 and 7 interspersed scans where b=0 s/mm2. The duration of DTI scans was approximately 11 minutes. The imaging volume was prescribed in axial orientation covering the entire cerebrum with the topmost slice just superior to the apex of the brain (Satterthwaite et al. 2014a). In addition to the DTI scan, a map of the main magnetic field (i.e., B0) was derived from a double-echo, gradient-recalled echo (GRE) sequence, allowing us to estimate field distortions in each dataset. Data Preprocessing Two consecutive 32-direction acquisitions were merged into a single 64-direction time- series. The skull was removed for each subject by registering a binary mask of a standard fractional anisotropy (FA) map (FMRIB58 FA) to each subject's DTI image using a rigid- body transformation (Smith et al., 2002). Eddy currents and subject motion were esti- mated and corrected using FSL's eddy tool (Andersson and Sotiropoulos 2016). Diffusion gradient vectors were then rotated to adjust for subject motion estimated by eddy. After the field map was estimated, distortion correction was applied to DTI data using FSL's FUGUE (Jenkinson et al., 2012). Lastly, DTI data was imported into DSI Studio software and the diffusion tensor was estimated at each voxel. DTI Tractography Whole-brain fiber tracking was implemented for each subject in DSI Studio using a modified fiber assessment by continuous tracking (FACT) algorithm with Euler interpola- tion, initiating 1,000,000 streamlines after removing all streamlines with length less than 10mm or greater than 400mm (Yeh et al., 2013). Fiber tracking was performed with an angular threshold of 45◦, a step size of 0.9375mm, and a fractional anisotropy (FA) thresh- old determined empirically by Otzu's method, which optimizes the contrast between fore- ground and background (Yeh et. al., 2013). Diffusivity measures (e.g., FA, mean diffusiv- ity, radial diffusivity, axial diffusivity) were calculated along the path of each reconstructed streamline. For each subject, tractography served as the basis for constructing structural brain networks. 33 Network Construction Following T1 reconstruction in FreeSurfer (version 5.3), cortical and subcortical gray matter was parcellated according to the Lausanne atlas (Cammoun et al., 2012), which includes whole-brain parcellations at multiple spatial scales (83, 129, 234, 463, and 1015 regions). Parcellations were defined in native space and co-registered to the first b = 0 volume of each subject's diffusion image using a rigid-body transform. To extend gray matter region labels beyond the gray-white boundary, the atlas labels were dilated by 4mm (Gu et al., 2015). Dilation involved filling non-labelled voxels with the statistical mode of neighboring labels. 234 dilated brain regions defined the nodes for each subject's structural brain network, which was represented as a weighted adjacency matrix A. Edges were defined where at least one streamline connected a pair of nodes end-to-end. Edge weights were primarily defined by the average FA along streamlines connecting any pair of nodes (Misic et al., 2016; Bohlken et al., 2016). See Figure 2. Functional Module Assignment For the 234- and 463-region parcellations, we calculated a purity index for each Lau- sanne label and corresponding voxels in the standard 7-system template image provided by Yeo et al. (2011). This measure quantifies the maximum overlap of cortical Lausanne labels and functional systems defined by Yeo et al. (2011). Each cortical Lausanne label was assigned to a functional system by calculating the non-zero mode of all voxels in each brain region. Subcortical regions were assigned to an eighth, subcortical module. The primary modular partition defined for 234-node networks is shown in Figure 2. To de- termine whether the functionally-defined network partition significantly fit the structural connectivity data beyond chance, we quantified the modularity quality index (formally defined below) of the functional partition imposed on structural brain networks. Briefly, the modularity quality of a network partition quantifies how well that partition maximizes the strength of within-module connections relative to a specified null model. Higher Q values indicate that modules are highly segregated within a network, with strong within- module connectivity and relatively weak between-module connectivity. We performed a permutation test to examine the significance of the modularity quality of the functional partition (QYeo) imposed on structural connectivity matrices. First, we permuted the as- signment of N nodes to functional modules 1000 times, preserving the number of nodes originally assigned to each module. We then calculated the modularity quality Qperm of randomly-defined network partitions imposed on each subject's connectivity matrix, build- ing a null distribution for Qperm. We used the calculated mean (µQperm) and standard de- viation (σQperm) of the null distribution to derive a z-score based on the observed QYeo for each subject (z-score = (QYeo−µQperm ) ). Finally, we calculated the mean z-score across all subjects to assess the significance of QYeo. σQperm 34 Measures of Modular Segregation We calculated the participation coefficient to quantify the relative balance of between- module versus within-module connectivity for each brain region. Intuitively, this measure describes the degree to which a brain region integrates information across distinct modules, or the degree to which a brain region shows provincial connectivity among regions in its own module. We define the participation coefficient Pi of node i as (cid:16)ki(m) (cid:17)2 ki Pi = 1− ∑ m∈M , (1) where m is a module in a set of modules M, and ki(m) is the weight of structural connec- tions between node i and all nodes in module m (Guimera and Amaral 2005; Rubinov and Sporns 2010). Moreover, Pi close to 1 indicates that a brain region is highly integrated with regions in other modules, while a Pi close to 0 indicates that a brain region is highly segregated, with strong connectivity among other regions in its own module. To quantify the segregation of specific modules, we average Pi across all brain regions assigned to the same module. To quantify global network segregation, we average Pi across all nodes in the network. Alternative Measures of Modular Segregation To ensure that our results were not dependent on specific network metrics, we cal- culated alternative measures of modular segregation. First, we calculated the average strength of all within-module connections (a measure of structural coherence), and the average strength of all between-module connections (a measure of structural integration) in the network (Gu et al., 2015). These metrics provide additional insights into the segre- gation of information processing within distinct modules, and the degree to which modules are integrated across the network (see Figure 4). Alternatively, we calculated the subject- specific modularity quality (Q) of group-level functional and structural network partitions. As discussed above, this measure provides an index of how well a network can be decom- posed into a hard partition where nodes within the same module demonstrate particularly strong connectivity beyond chance. We also calculated Qsub j for subject-specific consen- sus partitions (see detailed procedure below), which was not dependent on a group-level partition. We calculated the modularity Q of a network partition S based on the following modularity quality function: (cid:104) (cid:105) Q(S) = 1 2m ∑ i j Ai j − γPi j δ (gi,g j), (2) where m is the total weight of A, P represents the expected strength of connections accord- ing to a specified null model (Newman, 2004), γ is a structural resolution parameter that 35 determines the size of modules, and δ (gi,g j) is equal to unity when brain regions i and j are assigned to same community gi, and is zero otherwise. Community Detection in Structural Brain Networks Primary analyses relied on an a priori functional partition to define network modules. We additionally estimated network modules directly from the structural connectivity data using community detection procedures. Communities were defined by maximizing the modularity quality function using a generalization of the Louvain heuristic (Blondel et al., 2008; Mucha et al., 2010). Because the Louvain algorithm is degenerate (Good et al., 2010; Sporns and Betzel 2016), it is essential to perform modularity maximization multiple times in order to identify a stable consensus partition that accurately reflects the solutions offered by each optimization. Accordingly, we applied a locally greedy Louvain- like modularity-optimization procedure (Blondel et al., 2008) 100 times for each subject in order to define an "agreement" matrix A(cid:48) where A(cid:48) i j was equal to the probability that nodes i and j were assigned to the same community over the 100 iterations. If A(cid:48) was de- terministic (edge weights were binary), then the algorithm had converged and the resultant partition was defined as the consensus. Otherwise, we performed 100 iterations of modu- larity optimization on A(cid:48) in order to generate a new agreement matrix A(cid:48)(cid:48). This procedure was repeated until convergence (Lancichinetti and Fortunato, 2014). When performing modularity optimization on an agreement matrix (e.g., A(cid:48) or A(cid:48)(cid:48)), we defined an alternative null model P(cid:48) by permuting community assignments across nodes (Bassett et al., 2013). Once a consensus partition was identified for each subject, we computed a group- level consensus across the full PNC cohort (n=882). To do this, we used a Louvain-like procedure to detect communities in a group-level agreement matrix A(cid:48) group. Edge weights in A(cid:48) group were equal to the proportion of times that each pair of nodes was assigned to the same community across subject-level consensus partitions. As above, 100 iterations of modularity optimization were performed on A(cid:48) group became binary, indicating that the algorithm had converged on a group-level consensus partition. Both subject-level and group-level consensus partitions were computed over a wide range of γ ([0,4], in increments of 0.05) to explore variations in community structure. We plotted the number of group-level consensus modules as a function of γ, and found several plateaus indicating partition stability (Fenn et al., 2009; see Figure S1). In order to directly compare the organization of data-driven, modularity-based partitions and the a priori functional partition, we quantified the partition similarity using the z-score of the Rand coefficient (Traud et al., 2011). For two partitions X and Y , we calculated the Rand z-score in terms of the total number of node pairs in the network M, the number of pairs MX assigned to the same module in partition X, the number of pairs MY that are in the same module in partition Y , and the number of pairs of nodes wXY that are assigned to the same module group until the resulting A(cid:48)(cid:48) 36 both in partition X and in partition Y . The z-score of the Rand coefficient is defined by: zXY = 1 σwXY wXY − MXMY M , (3) where σwXY is the standard deviation of wXY . The mean partition similarity is determined by the mean value of zXY over all possible partition pairs for X (cid:54)= Y . Moreover, zXY denotes the similarity of partitions X and Y beyond chance. Figure S1 shows the similarity between all group-level structural partitions and the primary functional partition used in this study. Measures of Network Integration For each subject's structural brain network A, the topological length or distance of each edge Ai j was computed as the reciprocal of the edge weight ( 1 ). The path length between Ai j any pair of nodes is defined as the sum of the edge lengths along the shortest path connect- ing them (Rubinov and Sporns, 2010). Global efficiency provides a theoretical prediction of how easily information can flow across a network via the shortest path between all pairs of nodes, and is defined by (cid:18) (cid:19)−1 Eglob(G) = 1 n ∑ i∈N ∑ j∈N, j(cid:54)=i di j n− 1 , (4) where n is the number of nodes, and di j is the shortest path length between node i and node j. To examine the possible role of specific edges as integrative hub connections within the network, we calculated the weighted edge betweenness centrality (EBC) for each edge. Edge betweenness identifies important hub connections by providing a measure of how much a given connection participates in the shortest paths of communication through a network, and thus contributes to global efficiency (Brandes, 2001). EBC = ∑ hk ρi j hk ρhk , (5) where ρi j hk denotes the number of shortest paths between nodes h and k that include edge i j, and ρhk denotes the total number of shortest paths between h and k. After cal- culating EBC individually for each weighted network Ai j (n=882), we normalized each subjects(cid:48) EBC values by their maximum observed EBC, resulting in a bounded measure [0,1] (Gong et al., 2009). We calculated the mean normalized EBC for each network edge across subjects, and defined hub edges as those connections within the top quartile of 37 normalized edge betweenness across all network edges. Following group-level analysis, which identified a subset of edges that significantly strengthened with age, we performed a permutation-based test to assess whether connections that significantly strengthened with age were enriched for hub edges (see below). Group-level analyses Prior work has demonstrated that brain development is not a linear process (Paus et al., 1999, Shaw et al., 2006). Accordingly, group-level analyses of structural brain network metrics were flexibly modeled using penalized splines within a General Additive Model (GAM) implemented in the R package "mgcv" (https://cran.r-project.org/web/packages/mgcv/index.html; Wood 2004; Wood 2011). Such an approach allows for detection of nonlinearities in the relationship between age and measures of modular segregation without defining a set of functions a priori (such as polynomials). Importantly, the GAM estimates nonlinearities using restricted maximum likelihood (REML), and determines a penalty with increasing nonlinearity in order to avoid overfitting the data. Due to this penalty, the GAM only models nonlinearities when they explain additional variance in the data above and beyond linear effects. First, we used penalized splines to estimate nonlinear developmental patterns of mod- ular segregation. Within this model we included covariates for sex, head motion, and total network strength. Accordingly, the final model equations for estimating age effects on modular segregation (mean participation coefficient) were as follows: Modular segregation = spline(age) + sex + motion + total network strength An identical model was used when estimating age effects on the participation coefficient of individual brain regions. Similarly, we applied this model across all network edges in order to assess linear and nonlinear age effects on the strength of individual connections. For all analyses, multiple comparisons were controlled using the False Discovery Rate (q<0.05). Permutation Testing We performed permutation-based tests across network edges in order to assess (i) whether the edges that significantly strengthened with age were localized to within-module connections beyond chance, (ii) whether edges that significantly strengthen with age were enriched for hub edges, and (iii) whether these ages had elevated edge betweenness cen- trality beyond chance. First, we permuted a binary edge label specifying whether each edge connects nodes within or between modules 1000 times. Then for permuted samples of within- and between- module edges, we counted the number of edges that were shown to significantly strengthen 38 with age in group-level analysis. We then rank-ordered the number of edges shown to sig- nificantly strengthen with age for permuted within-module edge samples, and determined where the observed number of within-module edges that strengthen with age falls relative to this null distribution. Second, we evaluated whether edges that significantly strengthen with age were en- riched for hub edges. We permuted a binary edge label defining hub or non-hub edges 1000 times. For each permuted sample, we counted the number of edges that significantly strengthened with age in group-level analysis. Then, we rank-ordered the number of per- muted hub edges shown to significantly strengthen with age, and compared these values with the observed number of hub edges that strengthened with age. Third, we evaluated whether edges that significantly strengthen with age had higher edge betweenness centrality than anticipated by chance. We permuted normalized edge betweenness centrality values 1000 times. For each permuted sample, we calculated the mean EBC of within-module edges and between-module edges that significantly strength- ened with age. We rank-ordered the mean EBC of permuted within-module and between- module edges that strengthened with age, and compared these values with the observed means for within- and between-module edges separately (Figure 6D). Methodological Replications To verify that observed age-related increases in modular segregation were not simply due to specific network construction choices, we repeated developmental inferences on modular segregation using a variety of other parameters. First, we examined age effects on modular segregation (mean participation coefficient) using a data-driven structural par- tition identified at the group level (see Figure S2B., Figure 5B, and detailed procedure above). Alternatively, we also calculated the modularity quality index for each subject(cid:48)s optimal partition at γ=2.5 (QSub j), where a higher QSub j indicates greater modular sege- gration (Figure 5C). Next, we examined modular segregation (mean participation coeffi- cient) using the a priori functional partition assigned to a higher-resolution parcellation of the brain (463 nodes instead of 234; see Figure 5D). We also measured modular seg- regation of the functional partition using structural networks with alternative edge weight definitions. While primary analyses focused on FA-weighted structural networks, we also measured modular segregation in streamline-weighted networks (see Figure 5E), where edge weights were equal to the number of streamlines connecting a pair of nodes (Bassett et al., 2011), and additionally, where edge weights were defined by streamline density: the number of connecting streamlines divided by the total regional volume of each node pair (Baker et al., 2015; see Figure 5F). In addition to examining age-related patterns of modular segregation using alternative network measures and parameters, we also repeated analyses including the following additional covariates in the GAM described above: race, 39 maternal education, handedness, and total brain volume. Relationship Between Modular Segregation and Global Network Efficiency First, we examined age-related effects on global efficiency using the same GAM as above: Global efficiency = spline(age) + sex + motion + total network strength (see Figure 6A). The relationship between global efficiency and modular segregation was assessed within a GAM while controlling for age in addition to other covariates described above (Figure 6B). Moreover, the model equation was as follows: Modular segregation = Global efficiency + spline(age) + sex + motion + total network strength To assess whether global efficiency was related to the weight of specific network connec- tions that strengthened with age, we estimated the following GAMs: Global efficiency = Average strength of within-module edges + spline(age) + sex + motion + total network strength Global efficiency = Average strength of between-module edges + spline(age) + sex + mo- tion + total network strength (see Figure 6E and Figure 6F). Associations with Executive Function To examine the association between modular segregation and executive efficiency, we included a spline age term in the model to account for the variance associated with linear and nonlinear age-related changes in executive ability. The final model equation was as follows: Modular segregation = spline(age) + executive efficiency + sex + motion + total network strength Using the same GAM, we also evaluated the association between the segregation of indi- vidual modules (e.g., frontoparietal) and three cognitive efficiency factor scores: executive function, memory, and social cognition (see Figure 7A). We note that 2 participants of the full 882 sample had incomplete cognitive datasets: subsequent analyses examining associ- ations between executive function and modular segregation focused on the remaining 880 participants. Visualization of GAM model fits were created using the "visreg" package in R (https://cran.r-project.org/web/packages/visreg/). In Figure 3A, Figure 5, and Figure 40 6B, one outlying datapoint was beyond the axis range, and was excluded for visualiza- tion purposes only: group-level analyses and reported results include data points for all subjects. Mediation analyses Linear mediation analyses investigated whether age-related improvement in execu- tive function was mediated by modular segregation and/or global efficiency (Preacher and Hayes, 2008). First, we regressed out the effects of nuisance covariates (sex, head motion, and total network strength) on the independent (X), dependent (Y), and mediating (M) variables. The residuals were then used in our mediation analysis. The significance of the indirect effect was evaluated using bootstrapped confidence intervals within the R package "lavaan" (https://cran.r-project.org/web/packages/lavaan/). Specifically, we examined the total effect of age on executive performance (c path; Figure 7B), the relationship between age and modular segregation (a path), the relationship between modular segregation and executive function (b path), and the direct effect of age on executive efficiency after in- cluding modular segregation as a mediator in the model (c(cid:48) path). The significance of the indirect effect of age on executive function through the proposed mediator (modular seg- regation) was tested using bootstrapping procedures, which minimize assumptions about the sampling distribution (Preacher and Hayes, 2008). This approach involves calculating unstandardized indirect effects for each of 10,000 bootstrapped samples and calculating the 95% confidence interval. This procedure was repeated to assess (i) whether the seg- regation of the frontoparietal module mediated developmental improvements in executive function, (ii) whether the segregation of the default mode module mediated developmental improvements in social cognition, and (iii) whether age-related increases in global effi- ciency mediated improvements in executive function. Data Visualization Network partitions and regional results (Figure 2, Figure 3C, and Figure S2) were vi- sualized on the cortical white matter surface using FreeSurfer visualization tools in MAT- LAB. While age effects on the participation coefficient for subcortical brain regions are not visualized in Figure 3C, these regions were included in all analyses. Brain network visualizations in Figure 4 and Figure 6 were generated using BrainNet Viewer (Xia et al. 2013). 41 Supplemental References 1. Andersson, J.L. and Sotiropoulos, S.N. (2016). An integrated approach to correction for off-resonance effects and subject movement in diffusion MR imaging. Neuroim- age, 125,1063-1078. 2. Bassett, D.S., Porter, M.A., Wymbs, N.F., Grafton, S.T., Carlson, J.M. and Mucha, P.J. (2013). Robust detection of dynamic community structure in networks. Chaos, 23, 013142. 3. Fenn, D.J. Porter, M.A., McDonald, M., Williams, S., Johnson, N.F., and Jones, N.S. (2009). Dynamic communities in multichannel data: An application to the foreign exchange market during the 2007-2008 credit crisis. Chaos 19, 033119. 4. Giedd, J.N., Blumenthal, J., Jeffries, N.O., Castellanos, F.X., Liu, H., Zijdenbos, A., Paus, T., Evans, A.C., Rapoport, J.L. (1999). Brain development during childhood and adolescence: a longitudinal MRI study. Nature Neuroscience 2(10), 861-3. 5. Gong, G., He, Y., Concha, L., Lebel, C., Gross, D.W., Evans, A.C. and Beaulieu, C. (2009). Mapping anatomical connectivity patterns of human cerebral cortex using in vivo diffusion tensor imaging tractography. Cerebral Cortex, 19, 524-536. 6. Good, B. H., Montjoye, Y., and Clauset, A. (2010). Performance of modularity maximization in practical contexts. Physical Review E 81, 046106. 7. Gu, S., Pasqualetti, F., Cieslak, M., Telesford, Q.K., Alfred, B.Y., Kahn, A.E., Medaglia, J.D., Vettel, J.M., Miller, M.B., Grafton, S.T. and Bassett, D.S. (2015). Controllability of structural brain networks. Nature Communications, 6. 8. Lancichinetti, A. and Fortunato, S., 2012. Consensus clustering in complex net- works. Scientific reports, 2. 9. Newman, M.E.J., 2004. Analysis of weighted networks. Physical Review E, 70, 056131. 10. Paus, T., Zijdenbos, A., Worsley, K., Collins, D.L., Blumenthal, J., Giedd, J.N., Rapoport, J.L. and Evans, A.C. (1999). Structural maturation of neural pathways in children and adolescents: in vivo study. Science, 283, 1908-1911. 11. Shaw, P., Greenstein, D., Lerch, J., Clasen, L., Lenroot, R., Gogtay, N.E.E.A., Evans, A., Rapoport, J. and Giedd, J. (2006). Intellectual ability and cortical de- velopment in children and adolescents. Nature, 440, 676-679. 12. Xia, M., Wang, J. and He, Y. (2013). BrainNet Viewer: a network visualization tool for human brain connectomics. PLoS One, 8, e68910. 42
1904.10396
1
1904
2019-04-18T14:58:52
Is coding a relevant metaphor for building AI? A commentary on "Is coding a relevant metaphor for the brain?", by Romain Brette
[ "q-bio.NC", "cs.AI", "cs.LG" ]
Brette contends that the neural coding metaphor is an invalid basis for theories of what the brain does. Here, we argue that it is an insufficient guide for building an artificial intelligence that learns to accomplish short- and long-term goals in a complex, changing environment.
q-bio.NC
q-bio
Is coding a relevant metaphor for building AI? A commentary on "Is coding a relevant metaphor for the brain?", by Romain Brette Adam Santoro∗, Felix Hill∗, David G. T. Barrett, David Raposo, Matthew Botvinick, & Timothy Lillicrap {adamsantoro,felixhill,barrettdavid,draposo,botvinick,countzero}@google.com Deepmind, London UK Abstract Brette contends that the neural coding metaphor is an invalid basis for theories of what the brain does (Brette, 2019). Here, we argue that it is an insufficient guide for building an artificial intelligence (AI) that learns to accomplish short- and long-term goals in a complex, changing environment. The goal of neuroscience is to explain how the brain enables intelligent behaviour, while the goal of agent-based AI is to build agents that behave intelligently. Neuroscience, Brette attests, has suffered from an exaggerated (and technically inaccurate) concern for the codes transmitted by particular parts of the brain. In AI, on the other hand, some of the most notable recent progress has been made not by deeply considering neural coding and its implications, but by focusing on higher-level principles from optimization, learning, and control. Thanks to deep artificial networks trained via backpropagation, we now have artificial learning systems capable of impressive exhibitions of specific human-like skills, such as object recognition and language translation (e.g., He et al, 2016; Vaswani et al, 2017). In artificial, rather than biological, neural networks, we can more tractably characterize the relationship between a models neural codes, behaviour, and its external world. AI researchers have full access to their models' input data distribution, can visualise weights and activations in any part of the network and even make causal interventions on them, and can quickly implement new models informed by any coding hypotheses they may have. Nevertheless, in-depth analysis of a models internal representations is of increasingly rare concern for getting these models to work. Consider AlphaGo, which is one of the more ∗Equal Contribution 1 compelling recent breakthroughs in AI (Silver et al, 2016). Researchers on this project precisely defined the models goals, the dynamics underlying the models interactions with its environment, how the model plans its actions, and how the model learns. Each of these components contributes to the models success, and yet none of them fundamentally depend on considerations from neural coding. This is not to say that we cannot usefully apply representational analyses to such agents post-hoc, regardless of whether the representations satisfy Brettes criteria for neural codes (Barack et al, 2019). Indeed, since the earliest days of connectionism researchers have been interested in the neural codes that emerge when a clearly-specified learning algorithm is applied to a well-understood model trained to execute a particular task. A more recent and important collaboration between AI and neuroscience revealed insight into the conditions under which well-known codes can emerge: grid-cells can increasingly be understood as the product of particular optimization processes (Banino et al, 2018; Cueva et al, 2018). A key feature of these examples, however, is the central descriptive role given to the learning algorithms, architectures, and optimization objectives; neural coding was incidental, and in many cases the codes were not fundamental, privileged primitives on top of which the models were built (Marblestone et al, 2016). If the broader aim of agent-based AI (Russell et al, 2016) is to produce a system that accomplishes short- and long-term goals in a complex, changing environment, then there may be a more pernicious problem to the neural coding framework than it simply being out of vogue in modern AI. How internal responses arrive from given stimuli -- a goal that is implicit in the neural coding metaphor -- may be logically insufficient for producing intelligent behaviour. In outlining the reasons why, we recall arguments that any system -- artificial or biological -- needs to exert control over its environment to achieve intelligent behaviour . First, the observations with which an agent may compute do not exist as a prespecified dataset, independently of the agent's actions in the world. Rather, it is precisely the decisions that the agent takes in that world that determine the sensory data from which it learns. Second, [w]ithout an ongoing participation and perception of the world there is no meaning for an agent (Brooks, 1991). An agent participating in an external world that responds to its decisions learns useful, reliable, and meaningful interactions (Cisek, 1999). It is these meaningful interactions that ground the agents representations and allow them to be used for understanding and reasoning about its world. Therefore, insofar as neural coding is understood as a framework to help understand a systems internal stimulus-response patterns, it is a logically insufficient framework for designing AI because of its failure to engage with the agent-environment causal loop. Given these considerations, what, then, can we say about neural codings role in describing the brain? In neuroscience, we ultimately care about understanding how the brain enables intelligent behaviour. It is often argued that such an understanding cannot come from analyzing low-level, mechanistic details such as neural codes, because [a] description of neural activity and connections is not synonymous with knowing what they are doing to cause behavior. (Krakauer, 2017). For this level of understanding we need high-level computational and algorithmic theories that embrace agent-environment interactions. The history of AI tells us that the most useful principles, and the richest theoretical insights, emerged from studying 2 control, optimization, and learning processes rather than the particularities of representations or codes (Sutton, 2019). A focus on inferring such processes using our increasing quantities of neural data, rather than characterizing neural codes for their own sake, may also be the most productive way of making progress on understanding intelligent behaviour in humans and animals. Acknowledgments Thanks to Francis Song and Drew Jaegle for helpful comments. References Banino, A., Barry, C., Uria, B., Blundell, C., Lillicrap, T., Mirowski, P., ... & Wayne, G. (2018). Vector-based navigation using grid-like representations in artificial agents. Nature, 557(7705), 429. Barack, D. and Jaegle, A. (2019) Codes, Functions, and Causes: A Critique of Brettes Conceptual Analysis of Coding Brooks, R. A. (1991). Intelligence without reason. Cisek, P. (1999). Beyond the computer metaphor: Behaviour as interaction. Journal of Consciousness Studies, 6(11-12), 125-142. Cueva, C. J., & Wei, X. X. (2018). Emergence of grid-like representations by training recurrent neural networks to perform spatial localization. arXiv preprint arXiv:1803.07770. He, K., Zhang, X., Ren, S., & Sun, J. (2016). Deep residual learning for image recognition. In Proceedings of the IEEE conference on computer vision and pattern recognition (pp. 770-778). Krakauer, J. W., Ghazanfar, A. A., Gomez-Marin, A., MacIver, M. A., & Poeppel, D. (2017). Neuroscience needs behavior: correcting a reductionist bias. Neuron, 93(3), 480-490. Marblestone, A. H., Wayne, G., & Kording, K. P. (2016). Toward an integration of deep learning and neuroscience. Frontiers in computational neuroscience, 10, 94. Silver, D., Huang, A., Maddison, C. J., Guez, A., Sifre, L., Van Den Driessche, G., ... & Dieleman, S. (2016). Mastering the game of Go with deep neural networks and tree search. nature, 529(7587), 484. Russell, S. J., & Norvig, P. (2016). Artificial intelligence: a modern approach. Malaysia; Pearson Education Limited. Sutton, R. (2019). The Bitter Lesson. Available at http://incompleteideas.net/IncIdeas/ 3 BitterLesson.html Vaswani, A., Shazeer, N., Parmar, N., Uszkoreit, J., Jones, L., Gomez, A. N., ... & Polosukhin, I. (2017). Attention is all you need. In Advances in neural information processing systems (pp. 5998-6008). 4
1509.01972
1
1509
2015-09-07T10:21:42
The Brain Uses Reliability of Stimulus Information when Making Perceptual Decisions
[ "q-bio.NC" ]
In simple perceptual decisions the brain has to identify a stimulus based on noisy sensory samples from the stimulus. Basic statistical considerations state that the reliability of the stimulus information, i.e., the amount of noise in the samples, should be taken into account when the decision is made. However, for perceptual decision making experiments it has been questioned whether the brain indeed uses the reliability for making decisions when confronted with unpredictable changes in stimulus reliability. We here show that even the basic drift diffusion model, which has frequently been used to explain experimental findings in perceptual decision making, implicitly relies on estimates of stimulus reliability. We then show that only those variants of the drift diffusion model which allow stimulus-specific reliabilities are consistent with neurophysiological findings. Our analysis suggests that the brain estimates the reliability of the stimulus on a short time scale of at most a few hundred milliseconds.
q-bio.NC
q-bio
The Brain Uses Reliability of Stimulus Information when Making Perceptual Decisions Sebastian Bitzer1 [email protected] Stefan J. Kiebel1 [email protected] 1Department of Psychology, Technische Universitat Dresden, 01062 Dresden, Germany Abstract In simple perceptual decisions the brain has to identify a stimulus based on noisy sensory samples from the stimulus. Basic statistical considerations state that the reliability of the stimulus information, i.e., the amount of noise in the samples, should be taken into account when the decision is made. However, for perceptual decision making experiments it has been questioned whether the brain indeed uses the reliability for making decisions when confronted with unpredictable changes in stimulus reliability. We here show that even the basic drift diffusion model, which has frequently been used to explain experimental findings in perceptual decision making, implicitly relies on estimates of stimulus reliability. We then show that only those variants of the drift diffusion model which allow stimulus- specific reliabilities are consistent with neurophysiological findings. Our analysis suggests that the brain estimates the reliability of the stimulus on a short time scale of at most a few hundred milliseconds. 1 Introduction In perceptual decision making participants have to identify a noisy stimulus. In typical experiments, only two possibilities are considered [1]. The amount of noise on the stimulus is usually varied to manipulate task difficulty. With higher noise, participants' decisions are slower and less accurate. Early psychology research established that biased random walk models explain the response distri- butions (choice and reaction time) of perceptual decision making experiments [2]. These models describe decision making as an accumulation of noisy evidence until a bound is reached and cor- respond, in discrete time, to sequential analysis [3] as developed in statistics [4]. More recently, electrophysiological experiments provided additional support for such bounded accumulation mod- els, see [1] for a review. There appears to be a general consensus that the brain implements the mechanisms required for bounded accumulation, although different models were proposed for how exactly this accumulation is employed by the brain [5, 6, 1, 7, 8]. An important assumption of all these models is that the brain provides the input to the accumulation, the so-called evidence, but the most established models actually do not define how this evidence is computed by the brain [3, 5, 9, 1]. In this contribution, we will show that addressing this question offers a new perspective on how exactly perceptual decision making may be performed by the brain. Probabilistic models provide a precise definition of evidence: Evidence is the likelihood of a de- cision alternative under a noisy measurement where the likelihood is defined through a generative model of the measurements under the hypothesis that the considered decision alternative is true. In particular, this generative model implements assumptions about the expected distribution of mea- surements. Therefore, the likelihood of a measurement is large when measurements are assumed, 1 by the decision maker, to be reliable and small otherwise. For modelling perceptual decision making experiments, the evidence input, which is assumed to be pre-computed by the brain, should simi- larly depend on the reliability of measurements as estimated by the brain. However, this has been disputed before, e.g. [10]. The argument is that typical experimental setups make the reliability of each trial unpredictable for the participant. Therefore, it was argued, the brain can have no correct estimate of the reliability. This issue has been addressed in a neurally inspired, probabilistic model based on probabilistic population codes (PPCs) [7]. The authors have shown that PPCs can imple- ment perceptual decision making without having to explicitly represent reliability in the decision process. This remarkable result has been obtained by making the comprehensible assumption that reliability has a multiplicative effect on the tuning curves of the neurons in the PPCs. Reliability, therefore, was implicitly represented in the tuning curves of model neurons which leaves the open question why tuning curves exhibit a multiplicative effect of reliability. In this paper we will consider this question from a more conceptual perspective under the interpre- tation that the multiplicative effect on tuning curves results from adaptation of internal estimates of measurement reliability in the brain. We show that even a simple, widely used bounded accu- mulation model, the drift diffusion model, is based on some estimate of measurement reliability. Using this result, we will analyse the results of a perceptual decision making experiment [11] and will show that the recorded behaviour together with neurophysiological findings strongly favours the hypothesis that the brain weights evidence using a current estimate of measurement reliability, even when reliability changes unpredictably across trials. This paper is organised as follows: We first introduce the notions of measurement, evidence and likelihood in the context of the experimentally well-established random dot motion (RDM) stimulus. We define these quantities formally by resorting to a simple probabilistic model which has been shown to be equivalent to the drift diffusion model [12, 13]. This, in turn, allows us to formulate three competing variants of the drift diffusion model that either do not use trial-dependent reliability (variant CONST), or do use trial-dependent reliability of measurements during decision making (variants DDM and DEPC, see below for definitions). Finally, using data of [11], we show that only variants DDM and DEPC, which use trial-dependent reliability, are consistent with previous findings about perceptual decision making in the brain. 2 Measurement, evidence and likelihood in the random dot motion stimulus The widely used random dot motion (RDM) stimulus consists of a set of randomly located dots shown within an invisible circle on a screen [14]. From one video frame to the next some of the dots move into one direction which is fixed within a trial of an experiment, i.e., a subset of the dots moves coherently in one direction. All other dots are randomly replaced within the circle. Although there are many variants of how exactly to present the dots [15], the main idea is that the coherently moving dots indicate a motion direction which participants have to decide upon. By varying the proportion of dots which move coherently, also called the 'coherence' of the stimulus, the difficulty of the task can be varied effectively. We will now consider what kind of evidence the brain can in principle extract from the RDM stim- ulus in a short time window, for example, from one video frame to the next, within a trial. For simplicity we call this time window 'time point' from here on, the idea being that evidence is ac- cumulated over different time points, as postulated by bounded accumulation models in perceptual decision making [3, 1]. At a single time point, the brain can measure motion directions from the dots in the RDM display. By construction, a proportion of measurable motion directions will be into one specific direction, but, through the random relocation of other dots, the RDM display will also contain motion in random directions. Therefore, the brain observes a distribution of motion directions at each time point. This distribution can be considered a 'measurement' of the RDM stimulus made by the brain. Due to the randomness of each time frame, this distribution varies across time points and the variation in the distribution reduces for increasing coherences. We have illustrated this using rose histograms in Fig. 1 for three different coherence levels. To compute the evidence for the decision whether the RDM stimulus contains predominantly motion to one of the two considered directions, e.g., left and right, the brain must check how strongly these 2 Figure 1: Illustration of possible motion direction distributions that the brain can measure from an RDM stimulus. Rows are different time points, columns are different coherences. The true, underlying motion direction was 'left', i.e., 180◦. For low coherence (e.g., 3.2%) the measured distribution is very variable across time points and may indicate the presence of many different motion directions at any given time point. As coherence increases (from 9% to 25.6%), the true, underlying motion direction will increasingly dominate measured motion directions simultaneously leading to decreased variation of the measured distribution across time points. directions are represented in the measured distribution, e.g., by estimating the proportion of motion towards left and right. We call these proportions evidence for left, eleft, and evidence for right, eright. As the measured distribution over motion directions may vary strongly across time points, the computed evidences for each single time point may be unreliable. Probabilistic approaches weight evidence by its reliability such that unreliable evidence is not over-interpreted. The question is: Does the brain perform this reliability-based computation as well? More formally, for a given coherence, c, does the brain weight evidence by an estimate of reliability that depends on c: l = e · r(c)1 and which we call 'likelihood', or does it ignore changing reliabilities and use a weighting unrelated to coherence: e(cid:48) = e · ¯r? 3 Bounded accumulation models Bounded accumulation models postulate that decisions are made based on a decision variable. In particular, this decision variable is driven towards the correct alternative and is perturbed by noise. A decision is made, when the decision variable reaches a specific value. In the drift diffusion model, these three components are represented by drift, diffusion and bound [3]. We will now relate the typical drift diffusion formalism to our notions of measurement, evidence and likelihood by linking the drift diffusion model to probabilistic formulations. In the drift diffusion model, the decision variable evolves according to a simple Wiener process with drift. In discrete time the change in the decision variable y can be written as δy = yt − yt−δt = vδt + √ (1) where v is the drift, t ∼ N (0, 1) is Gaussian noise and s controls the amount of diffusion. This equation bears an interesting link to how the brain may compute the evidence. For example, it has been stated in the context of an experiment with RDM stimuli with two decision alternatives that the change in y, often called 'momentary evidence', "is thought to be a difference in firing rates of direction selective neurons with opposite direction preferences." [11, Supp. Fig. 6] Formally: δtst δy = ρleft,t − ρright,t (2) 1For convenience, we use imprecise denominations here. As will become clear below, l is in our case a Gaussian log-likelihood, hence, the linear weighting of evidence by reliability. 3 0°45°90°135°180°225°270°315°246810120°45°90°135°180°225°270°315°24681012140°45°90°135°180°225°270°315°510152025300°45°90°135°180°225°270°315°1234567890°45°90°135°180°225°270°315°2468100°45°90°135°180°225°270°315°5101520253035 3.2%time point 1 9.0%25.6%time point 2 √ l = − log( 2πδtσ) − 1 2 (xt ± 1)2 δtσ2 . (5) where ρleft,t is the firing rate of the population selective to motion towards left at time point t. Because the firing rates ρ depend on the considered decision alternative, they represent a form of evidence extracted from the stimulus measurement instead of the stimulus measurement itself (see our definitions in the previous section). It is unclear, however, whether the firing rates ρ just represent the evidence (ρ = e(cid:48)) or whether they represent the likelihood, ρ = l, i.e., the evidence weighted by coherence-dependent reliability. To clarify the relation between firing rates ρ, evidence e and likelihood l we consider probabilistic models of perceptual decision making. Several variants have been suggested and related to other forms of decision making [6, 16, 9, 7, 12, 17, 18]. For its simplicity, which is sufficient for our argument, we here consider the model presented in [13] for which a direct transformation from probabilistic model to the drift diffusion model has already been shown. This model defines two Gaussian generative models of measurements which are derived from the stimulus: p(xtleft) = N (−1, δtσ2) (3) where σ represents the variability of measurements expected by the brain. Similarly, it is assumed that the measurements xt are sampled from a Gaussian with variance σ2 which captures variance both from the stimulus and due to other noise sources in the brain: p(xtright) = N (1, δtσ2) (4) Evidence for a decision is computed in this model by calculating the likelihood of a measurement xt under the hypothesised generative models. To be precise we consider the log-likelihood which is xt ∼ N (±1, δtσ2). There are three important points: 1) The first term on the right hand side means that l increases independently of the stimulus xt for decreasing σ. This contribution cancels when the difference between the likelihoods for left and right is computed. 2) The likelihood is large for a measurement xt, when xt is close to the values hypothesised for the decision alternatives, i.e., −1 and 1. 3) The contribution of the stimulus is weighted by the assumed reliability r = σ−2. This model of the RDM stimulus is simple but captures the most important properties of the stim- ulus. In particular, a high coherence RDM stimulus has a large proportion of motion in the correct direction with very low variability of measurements whereas a low coherence RDM stimulus tends to have lower proportions of motion in the correct direction, with high variability (cf. Fig. 1). The Gaussian model captures these properties by adjusting the noise variance such that a high coherence corresponds to low noise and low coherence to high noise: Under high noise the values xt will vary strongly and tend to be rather distant from −1 and 1, whereas for low noise the values xt will be close to −1 or 1 with low variability. Hence, as expected, the model produces large evidences/likelihoods for low noise and small evidences/likelihoods for high noise. This intuitive relation between stimulus and probabilistic model is the basis for us to proceed to show that the reliability of the stimulus r, connected to the coherence level c, appears at a prominent position in the drift diffusion model. Crucially, the drift diffusion model can be derived as the sum of log-likelihood ratios across time [3, 9, 12, 13]. In particular, a discrete time drift diffusion process can be derived by subtracting the likelihoods of Eq. (5): (xt + 1)2 − (xt − 1)2 2rxt δt . δy = lright − lleft = = 2δtσ2 (6) Consequently, the change in y is Gaussian: δy ∼ N (2r/δt, 4r2σ2/δt). This replicates the model described in [11, Supp. Fig. 6] where the parameterisation of the model, however, more di- rectly followed that of the Gaussian distribution and did not explicitly take time into account: δy ∼ N (Kc, S2), where K and S are free parameters and c is coherence of the RDM stimulus. By analogy to the probabilistic model, we, therefore, see that the model in [11] implicitly assumes that reliability r depends on coherence c. More generally, the parameters of the drift diffusion model of Eq. (1) and that of the probabilistic model can be expressed as functions of each other [13]: v = ± 2 δt2 σ2 = ±r 2 δt2 4 (7) s = 2σ δtσ2 = r 2σ δt . (8) These equations state that both drift v and diffusion s depend on the assumed reliability r of the measurements x. Does the brain use and necessarily compute this reliability which depends on coherence? In the following section we answer this question by comparing how well three variants of the drift diffusion model, that implement different assumptions about r, conform to experimental findings. 4 Use of reliability in perceptual decision making: experimental evidence We first show that different assumptions about the reliability r translate to variants of the drift dif- fusion model. We then fit all variants to behavioural data (performances and mean reaction times) of an experiment for which neurophysiological data has also been reported [11] and demonstrate that only those variants which allow reliability to depend on coherence level lead to accumulation mechanisms which are consistent with the neurophysiological findings. 4.1 Drift diffusion model variants For the drift diffusion model of Eq. (1) the accuracy A and mean decision time T predicted by the model can be determined analytically [9]: A = 1 − 1 + exp( 2vb s2 ) 1 (cid:18) vb (cid:19) s2 T = b v tanh (9) (10) where b is the bound. These equations highlight an important caveat of the drift diffusion model: Only two of the three parameters can be determined uniquely from behavioural data. For fitting the model one of the parameters needs to be fixed. In most cases, the diffusion s is set to c = 0.1 arbitrarily [9], or is fit with a constant value across stimulus strengths [11]. We call this standard variant of the drift diffusion model the DDM. If s is constant across stimulus strengths, the other two parameters of the model must explain dif- ferences in behaviour, between stimulus strengths, by taking on values that depend on stimulus strength. Indeed, it has been found that primarily drift v explains such differences, see also be- low. Eq. (7) states that drift depends on estimated reliability r. So, if drift varies across stimulus strengths, this strongly suggests that r must vary across stimulus strengths, i.e., that r must depend on coherence: r(c). However, the drift diffusion formalism allows for two other obvious variants of parameterisation. One in which the bound b is constant across stimulus strengths, b = ¯b, and, conversely, one in which drift v is constant across stimulus strengths, v = ¯v ∝ ¯r (Eq. 7). We call these variants DEPC and CONST, respectively, for their property to weight evidence by reliability that either depends on coherence, r(c), or not, ¯r. 4.2 Experimental data In the following we will analyse the data presented in [11]. This data set has two major advantages for our purposes: 1) Reported accuracies and mean reaction times (Fig. 1d,f) are averages based on 15,937 trials in total. Therefore, noise in this data set is minimal (cf. small error bars in Fig. 1d,f) such that any potential effects of overfitting on found parameter values will be small, especially in relation to the effect induced by different stimulus strengths. 2) The behavioural data is accompanied by recordings of neurons which have been implicated in the decision making process. We can, therefore, compare the accumulation mechanisms resulting from the fit to behaviour with the actual neurophysiological recordings. Furthermore, the structure of the experiments was such that the stimulus in subsequent trials had random strength, i.e., the brain could not have estimated stimulus strength of a trial before the trial started. In the experiment of [11], that we consider here, two monkeys performed a two-alternative forced choice task based on the RDM stimulus. Data for eight different coherences were reported. To avoid 5 ceiling effects, which prevent the unique identification of parameter values in the drift diffusion model, we exclude those coherences which lead to an accuracy of 0.5 (random choices) or to an accuracy of 1 (perfect choices). The behavioural data of the remaining six coherence levels are presented in Table 1. Table 1: Behavioural data of [11] used in our analysis. RT = reaction time. coherence (%): accuracy (fraction): mean RT (ms): 3.2 0.63 613 6.4 0.76 590 9 0.79 580 12 0.89 535 25.6 0.99 440 The analysis of [11] revealed a nondecision time, i.e., a component of the reaction time that is unrelated to the decision process (cf. [3]) of ca. 200ms. Using this estimate, we determined the mean decision time T by subtracting 200ms from the mean reaction times shown in Table 1. The main findings for the neural recordings, which replicated previous findings [19, 1], were that i) firing rates at the end of decisions were similar and, particularly, showed no significant relation to coherence [11, Fig. 5] whereas ii) the buildup rate of neural firing within a trial had an approximately linear relation to coherence [11, Fig. 4]. 4.3 Fits of drift diffusion model variants to behaviour We can easily fit the model variants (DDM, DEPC and CONST) to accuracy A and mean decision time T using Eqs. (9) and (10). In accordance with previous approaches we selected values for the respective redundant parameters. Since the redundant parameter value, or its inverse, simply scales the fitted parameter values (cf. Eqs. 9 and 10), the exact value is irrelevant and we fix, in each model variant, the redundant parameter to 1. Figure 2: Fitting results: values of the free parameters, that replicate the accuracy and mean RT recorded in the experiment (Table 1), in relation to coherence. The remaining, non-free parameter was fixed to 1 for each variant. Left: the DDM variant with free parameters drift v (green) and bound b (purple). Middle: the DEPC variant with free parameters v and diffusion s (orange). Right: the CONST variant with free parameters s and b. Fig. 2 shows the inferred parameter values. In congruence with previous findings, the DDM variant explained variation in behaviour due to an increasing coherence mostly with an increasing drift v (green in Fig. 2). Specifically, drift and coherence appear to have a straightforward, linear relation. The same finding holds for the DEPC variant. In contrast to the DDM variant, however, which also exhibited a slight increase in the bound b (purple in Fig. 2) with increasing coherence, the DEPC variant explained the corresponding differences in behaviour by decreasing diffusion s (orange in Fig. 2). As the drift v was fixed in CONST, this variant explained coherence-dependent behaviour with large and almost identical changes in both diffusion s and bound b such that large parameter values occurred for small coherences and the relation between parameters and coherence appeared to be quadratic. We further investigated the properties of the model variants with the fitted parameter values. The top row of Fig. 3 shows example drift diffusion trajectories (y in Eq. (1)) simulated at a resolution of 1ms for two coherences. Following [11], we interpret y as the decision variables represented by the 6 051015202530coherence (%)0510152025bDDM051015202530coherence (%)0.000.010.020.030.040.05sDEPC051015202530coherence (%)01020304050607080sCONST0.000.020.040.060.080.10v0.0000.0010.0020.0030.004v02004006008001000120014001600b Figure 3: Drift-diffusion properties of fitted model variants. Top row: 15 example trajectories of y for different model variants with fitted parameters for 6.4% (blue) and 25.6% (yellow) coherence. Trajectories end when they reach the bound for the first time which corresponds to the decision time in that simulated trial. Notice that the same random samples of  were used across variants and coherences. Bottom row: Trajectories of y averaged over trials in which the first alternative (top bound) was chosen for the three model variants. Format of the plots follows that of [8, Supp. Fig. 4]: Left panels show the buildup of y from the start of decision making for the 5 different coherences. Right panels show the averaged drift diffusion trajectories when aligned to the time that a decision was made. firing rates of neurons in monkey area LIP. These plots exemplify that the DDM and DEPC variants lead to qualitatively very similar predictions of neural responses whereas the trajectories produced by the CONST variant stand out, because the neural responses to large coherences are predicted to be smaller than those to small coherences. We have summarised predicted neural responses to all coherences in the bottom row of Fig. 3 where we show averages of y across 5000 trials either aligned to the start of decision making (left pan- els) or aligned to the decision time (right panels). These plots illustrate that the DDM and DEPC variants replicate the main neurophysiological findings of [11]: Neural responses at the end of the decision were similar and independent of coherence. For the DEPC variant this was built into the model, because the bound was fixed. For the DDM variant the bound shows a small dependence on coherence, but the neural responses aligned to decision time were still very similar across coher- ences. The DDM and DEPC variants, further, replicate the finding that the buildup of neural firing depends approximately linear on coherence (normalised mean square error of a corresponding linear model was 0.04 and 0.03, respectively). In contrast, the CONST variant exhibited an inverse rela- tion between coherence and buildup of predicted neural response, i.e., buildup was larger for small coherences. Furthermore, neural responses at decision time strongly depended on coherence. There- fore, the CONST variant, as the only variant which does not use coherence-dependent reliability, is also the only variant which is clearly inconsistent with the neurophysiological findings. 5 Discussion We have investigated whether the brain uses online estimates of stimulus reliability when making simple perceptual decisions. From a probabilistic perspective fundamental considerations suggest that using accurate estimates of stimulus reliability lead to better decisions, but in the field of percep- tual decision making it has been questioned that the brain estimates stimulus reliability on the very short time scale of a few hundred milliseconds. By using a probabilistic formulation of the most 7 0100200300400500600700800time from start (ms)201001020yDDM 6.425.60100200300time from start (ms)05101520mean of y3.26.49.012.025.6-300-200-100DTtime from end (ms)0100200300400500600700800time from start (ms)1.00.50.00.51.0DEPC0100200300time from start (ms)0.00.20.40.60.81.0-300-200-100DTtime from end (ms)0100200300400500600700800time from start (ms)6004002000200400600CONST0100200300time from start (ms)0200400600800100012001400-300-200-100DTtime from end (ms) widely accepted model we were able to show that only those variants of the model which assume online reliability estimation are consistent with reported experimental findings. Our argument is based on a strict distinction between measurements, evidence and likelihood which may be briefly summarised as follows: Measurements are raw stimulus features that do not relate to the decision, evidence is a transformation of measurements into a decision relevant space reflecting the decision alternatives and likelihood is evidence scaled by a current estimate of measurement reliabilities. It is easy to overlook this distinction at the level of bounded accumulation models, such as the drift diffusion model, because these models assume a pre-computed form of evidence as input. However, this evidence has to be computed by the brain, as we have demonstrated based on the example of the RDM stimulus and using behavioural data. We chose one particular, simple probabilistic model, because this model has a direct equivalence with the drift diffusion model which was used to explain the data of [11] before. Other models may have not allowed conclusions about reliability estimates in the brain. In particular, [13] introduced an alternative model that also leads to equivalence with the drift diffusion model, but explains dif- ferences in behaviour by different mean measurements and their representations in the generative model. Instead of varying reliability across coherences, this model would vary the difference of means in the second summand of Eq. (5) directly without leading to any difference on the drift diffusion trajectories represented by y of Eq. (1) when compared to those of the probabilistic model chosen here. The interpretation of the alternative model of [13], however, is far removed from basic assumptions about the RDM stimulus: Whereas the alternative model assumes that the reliability of the stimulus is fixed across coherences, the noise in the RDM stimulus clearly depends on coherence. We, therefore, discarded the alternative model here. As a slight caveat, the neurophysiological findings, on which we based our conclusion, could have been the result of a search for neurons that exhibit the properties of the conventional drift diffusion model (the DDM variant). We cannot exclude this possibility completely, but given the wide range and persistence of consistent evidence for the standard bounded accumulation theory of decision making [1, 20] we find it rather unlikely that the results in [19] and [11] were purely found by chance. Even if our conclusion about the rapid estimation of reliability by the brain does not en- dure, our formal contribution holds: We clarified that the drift diffusion model in its most common variant (DDM) is consistent with, and even implicitly relies on, coherence-dependent estimates of measurement reliability. In the experiment of [11] coherences of the RDM stimulus were chosen randomly for each trial. Consequently, participants could not predict the reliability of the RDM stimulus for the upcoming trial, i.e., the participants' brains could not have had a good estimate of stimulus reliability at the start of a trial. Yet, our analysis strongly suggests that coherence-dependent reliabilities were used during decision making. The brain, therefore, must had adapted reliability within trials even on the short timescale of a few hundred milliseconds. On the level of analysis dictated by the drift diffusion model we cannot observe this adaptation. It only manifests itself as a change in mean drift that is assumed to be constant within a trial. First models of simultaneous decision making and reliability estimation have been suggested [21], but clearly more work in this direction is needed to elucidate the underlying mechanism used by the brain. References [1] Joshua I Gold and Michael N Shadlen. The neural basis of decision making. Annu Rev Neu- rosci, 30:535 -- 574, 2007. [2] I. D. John. A statistical decision theory of simple reaction time. Australian Journal of Psy- chology, 19(1):27 -- 34, 1967. [3] R. Duncan Luce. Response Times: Their Role in Inferring Elementary Mental Organization. Number 8 in Oxford Psychology Series. Oxford University Press, 1986. [4] Abraham Wald. Sequential Analysis. Wiley, New York, 1947. [5] Xiao-Jing Wang. Probabilistic decision making by slow reverberation in cortical circuits. Neu- ron, 36(5):955 -- 968, Dec 2002. [6] Rajesh P N Rao. Bayesian computation in recurrent neural circuits. Neural Comput, 16(1):1 -- 38, Jan 2004. 8 [7] Jeffrey M Beck, Wei Ji Ma, Roozbeh Kiani, Tim Hanks, Anne K Churchland, Jamie Roitman, Michael N Shadlen, Peter E Latham, and Alexandre Pouget. Probabilistic population codes for Bayesian decision making. Neuron, 60(6):1142 -- 1152, December 2008. [8] Anne K Churchland, R. Kiani, R. Chaudhuri, Xiao-Jing Wang, Alexandre Pouget, and M. N. Shadlen. Variance as a signature of neural computations during decision making. Neuron, 69(4):818 -- 831, Feb 2011. [9] Rafal Bogacz, Eric Brown, Jeff Moehlis, Philip Holmes, and Jonathan D. Cohen. The physics of optimal decision making: a formal analysis of models of performance in two-alternative forced-choice tasks. Psychol Rev, 113(4):700 -- 765, October 2006. [10] Michael N Shadlen, Roozbeh Kiani, Timothy D Hanks, and Anne K Churchland. Neurobi- ology of decision making: An intentional framework. In Christoph Engel and Wolf Singer, editors, Better Than Conscious? Decision Making, the Humand Mind, and Implications For Institutions. MIT Press, 2008. [11] Anne K Churchland, Roozbeh Kiani, and Michael N Shadlen. Decision-making with multiple alternatives. Nat Neurosci, 11(6):693 -- 702, Jun 2008. [12] Peter Dayan and Nathaniel D Daw. Decision theory, reinforcement learning, and the brain. Cogn Affect Behav Neurosci, 8(4):429 -- 453, Dec 2008. [13] Sebastian Bitzer, Hame Park, Felix Blankenburg, and Stefan J Kiebel. Perceptual decision making: Drift-diffusion model is equivalent to a bayesian model. Frontiers in Human Neuro- science, 8(102), 2014. [14] W. T. Newsome and E. B. Par´e. A selective impairment of motion perception following lesions of the middle temporal visual area MT. J Neurosci, 8(6):2201 -- 2211, June 1988. [15] Praveen K. Pilly and Aaron R. Seitz. What a difference a parameter makes: a psychophysical comparison of random dot motion algorithms. Vision Res, 49(13):1599 -- 1612, Jun 2009. [16] Angela J. Yu and Peter Dayan. Inference, attention, and decision in a Bayesian neural ar- chitecture. In Lawrence K. Saul, Yair Weiss, and L´eon Bottou, editors, Advances in Neural Information Processing Systems 17, pages 1577 -- 1584. MIT Press, Cambridge, MA, 2005. [17] Alec Solway and Matthew M. Botvinick. Goal-directed decision making as probabilistic in- ference: a computational framework and potential neural correlates. Psychol Rev, 119(1):120 -- 154, January 2012. [18] Yanping Huang, Abram Friesen, Timothy Hanks, Mike Shadlen, and Rajesh Rao. How prior probability influences decision making: A unifying probabilistic model. In P. Bartlett, F.C.N. Pereira, C.J.C. Burges, L. Bottou, and K.Q. Weinberger, editors, Advances in Neural Informa- tion Processing Systems 25, pages 1277 -- 1285. 2012. [19] Jamie D Roitman and Michael N Shadlen. Response of neurons in the lateral intraparietal area during a combined visual discrimination reaction time task. J Neurosci, 22(21):9475 -- 9489, Nov 2002. [20] Timothy D. Hanks, Charles D. Kopec, Bingni W. Brunton, Chunyu A. Duan, Jeffrey C. Erlich, and Carlos D. Brody. Distinct relationships of parietal and prefrontal cortices to evidence accumulation. Nature, Jan 2015. [21] Sophie Den`eve. Making decisions with unknown sensory reliability. Front Neurosci, 6:75, 2012. 9
1909.02931
1
1909
2019-09-06T14:31:54
Integrated Information in the Spiking-Bursting Stochastic Model
[ "q-bio.NC" ]
This study presents a comprehensive analytic description in terms of the empirical "whole minus sum" version of Integrated Information in comparison to the "decoder based" version for the "spiking-bursting" discrete-time, discrete-state stochastic model, which was recently introduced to describe a specific type of dynamics in a neuron-astrocyte network. The "whole minus sum" information may change sign, and an interpretation of this transition in terms of "net synergy" is available in the literature. This motivates our particular interest to the sign of the "whole minus sum" information in our analytical consideration. The behavior of the "whole minus sum" and "decoder based" information measures are found to bear a lot of similarity, showing their mutual asymptotic convergence as time-uncorrelated activity is increased, with the sign transition of the "whole minus sum" information associated to a rapid growth in the "decoder based" information. The study aims at creating a theoretical base for using the spiking-bursting model as a well understood reference point for applying Integrated Information concepts to systems exhibiting similar bursting behavior (in particular, to neuron-astrocyte networks). The model can also be of interest as a new discrete-state test bench for different formulations of Integrated Information.
q-bio.NC
q-bio
Integrated Information in the Spiking-Bursting Stochastic Model Oleg Kanakov,1 Susanna Gordleeva,1 and Alexey Zaikin1, 2, 3, ∗ 1Lobachevsky State University of Nizhny Novgorod, Nizhny Novgorod, Russia 2Institute for Women's Health and Department of Mathematics, 3Department of Pediatrics, Faculty of Pediatrics, Sechenov University, Moscow, Russia University College London, London, United Kingdom This study presents a comprehensive analytic description in terms of the empirical "whole mi- nus sum" version of Integrated Information in comparison to the "decoder based" version for the "spiking-bursting" discrete-time, discrete-state stochastic model, which was recently introduced to describe a specific type of dynamics in a neuron-astrocyte network. The "whole minus sum" in- formation may change sign, and an interpretation of this transition in terms of "net synergy" is available in the literature. This motivates our particular interest to the sign of the "whole minus sum" information in our analytical consideration. The behavior of the "whole minus sum" and "decoder based" information measures are found to bear a lot of similarity, showing their mutual asymptotic convergence as time-uncorrelated activity is increased, with the sign transition of the "whole minus sum" information associated to a rapid growth in the "decoder based" information. The study aims at creating a theoretical base for using the spiking-bursting model as a well un- derstood reference point for applying Integrated Information concepts to systems exhibiting similar bursting behavior (in particular, to neuron-astrocyte networks). The model can also be of interest as a new discrete-state test bench for different formulations of Integrated Information. I. INTRODUCTION Integrated information (II) [1 -- 4] is a measure of inter- nal information exchange in complex systems which was initially proposed to quantify consciousness [5]. This ini- tial aim still remaining a matter of research and debate [6 -- 9], the II concept itself is by now a widely acknowl- edged tool in the field of complex dynamics analysis [10 -- 12]. The general concept gave rise to specific "empirical" formalizations of II [13 -- 16] aimed at computability from empirical probability distributions based on real data. For a systematic taxonomy of II measures see [17], and a comparative study of empirical II measures in applica- tion to Gaussian autoregressive network models has been recently done in [18]. A recent study [19] addressed the role of astrocytic reg- ulation of neurotransmission [20] in generating positive II by small networks of brain cells -- neurons and astro- cytes. Empirical "whole minus sum" II as defined in [13] was calculated in [19] from the time series produced by a biologically realistic model of neuro-astrocytic networks. A simplified, analytically tractable stochastic "spiking- bursting" model (in complement to the realistic one) was designed to describe a specific type of activity in neuro- astrocytic networks which manifests itself as a sequence of intermittent system-wide excitations of rapid pulse trains ("bursts") on the background of random "spik- ing" activity in the network. The spiking-bursting model is a discrete-time, discrete-state stochastic process which mimics the main features of this behavior. The model was successfully used in [19] to produce semi-analytical estimates of II in good agreement with direct compu- ∗ E-mail: [email protected] tation of II from time series of the biologically realistic network model. The present study aims at creating a theoretical base for using the spiking-bursting model of [19] as a well un- derstood reference point for applying Integrated Infor- mation concepts to systems exhibiting similar bursting behavior (in particular, to other neuron-astrocyte net- works). We also aim at extending the knowledge of com- parative features of different empirical II measures, which are currently available mainly in application to Gaussian autoregressive models [17, 18], by applying two such mea- sures [13, 16] to our discrete-state model. In Sections II, III we specify the definitons of the used II measures and the model. Specific properties of the model which lead to a redundance in its parameter set are addressed in Section IV. In Section V we provide with an analytical treatment for the empirical "whole minus sum" [13] version of II in application to our model. This choice among other empirical II measures is inherited from the preceding study [19] and is in part due to its easy analyt- ical tractability, and also due to its ability to change sign, which naturally identifies a transition point in the param- eter space. This property may be considered a violation of the natural non-negativeness requirement for II [16]; on the other hand, the sign of the "whole minus sum" in- formation has been given interpretation in terms of "net synergy" [21] as a degree of redundancy in the evolution of a system [18]. In this sense this transition may be viewed as a useful marker in its own right in the toolset of measures for complex dynamics. This motivates our particular focus on identifying the sign transition of the "whole minus sum" information in the parameter space of the model. We also identify a scaling of II with a parameter determining time correlations of the bursting (astrocytic) subsystem when these correlations are weak. In Section VI we compare the outcome of the "whole minus sum" II measure [13] to the "decoder based" mea- sure Φ∗, which was specifically designed in [16] to sat- isfy the non-negativeness property. We compute Φ∗ di- rectly by definition from known probability distributions of the model. Despite their inherent difference consisting in changing or not changing sign, the two compared mea- sures are shown to bear similarities in their dependence upon model parameters, including the same scaling with the time correlation parameter. II. DEFINITION OF II MEASURES IN USE The empirical "whole minus sum" version of II is for- mulated according to [13] as follows. Consider a station- ary stochastic process ξ(t) (binary vector process), whose instantaneous state is described by N binary digits (bits), each identified with a node of the network (neuron). The full set of N nodes ("system") can be split into two non- overlapping non-empty subsets ("subsystems") A and B, such splitting further referred to as bipartition AB. De- note by x = ξ(t) and y = ξ(t+τ ) two states of the process separated by a specified time interval τ 6= 0. States of the subsystems are denoted as xA, xB, yA, yB. Mutual information between x and y is defined as where Ixy = Hx + Hy − Hxy, Hx = −Xx p(x) log2 p(x) (1) (2) is entropy (base 2 logarithm gives result in bits), summa- tion is hereinafter assumed to be taken over the whole range of the index variable (here x), Hy = Hx due to assumed stationarity. Next, a bipartition AB is considered, and "effective information" as a function of the particular bipartition is defined as Φeff(AB) = Ixy − IxA,yA − IxB ,yB . (3) II is then defined as effective information calculated for a specific bipartition ABMIB ("minimum information bipartition") which minimizes specifically normalized ef- fective information: Φ = Φeff(ABMIB), ABMIB = argminAB(cid:20) Φeff(AB) min{H(xA), H(xB)}(cid:21) . (4a) (4b) Note that this definition prohibits positive II, whenever Φeff turns out to be zero or negative for at least one bipartition AB. We compare the result of the "whole minus sum" effec- tive information (3) to the "decoder based" information measure Φ∗, which is modified from its original formula- tion of [16] by setting the logarithms base to 2 for con- sistency: Φ∗(AB) = Ixy − I ∗ xy(AB), (5a) where 2 I ∗ xy(AB) = max β "−Xy +Xxy p(y) log2Xx p(x)qAB(yx)β p(xy) log2 qAB(yx)β# , (5b) qAB(yx) = p(yAxA)p(yBxB) = p(xAyA)p(xByB) p(xA)p(xB) . (5c) III. SPIKING-BURSTING STOCHASTIC MODEL We consider a stochastic model, which produces a binary vector valued, discrete-time stochastic process. In keeping with [19], this "spiking-bursting" model is defined as a combination M = {V, S} of a time- correlated dichotomous component V which turns on and off system-wide bursting (that mimics global bursting of a neuronal network, when each neuron produces a train of pulses at a high rate [19]), and a time-uncorrelated component S describing spontaneous (spiking) activity (corresponding to a background random activity in a neural network characterized by relatively sparse ran- dom appearance of neuronal pulses -- spikes [19]) oc- curring in the absence of a burst. The model mimics the spiking-bursting type of activity which occurs in a neuro-astrocytic network, where the neural subsystem normally exhibits time-uncorrelated patterns of spiking activity, and all neurons are under the common influ- ence of the astrocytic subsystem, which is modeled by the dichotomous component V and sporadically induces simultaneous bursting in all neurons. A similar network architecture with a "master node" spreading its influence on subordinated nodes was considered for example in [1] (Figure 4b therein). The model is defined as follows. At each instance of (discrete) time the state of the dichotomous component can be either "bursting" with probability pb, or "sponta- neous" (or "spiking") with probability ps = 1−pb. While in the bursting mode, the instantaneous state of the re- sulting process x = ξ(t) is given by all ones: x = 11..1 (further abbreviated as x = 1). In case of spiking, the state x is a (time-uncorrelated) random variate described by a discrete probability distribution sx (where an occur- rence of '1' in any bit is referred to as a "spike"), so that the resulting one-time state probabilities read p(x 6= 1) = pssx, p(x = 1) = p1, p1 = pss1 + pb, (6a) (6b) where s1 is the probability of spontaneous occurrence of x = 1 (hereafter referred to as a system-wide simultane- ous [22] spike) in the absence of a burst. To describe two-time joint probabilities for x = ξ(t) and y = ξ(t + τ ), consider a joint state xy which is a con- catenation of bits in x and y. The spontaneous activity is assumed to be uncorrelated in time, which leads to the factorization sxy = sxsy. (7) The time correlations of the dichotomous component [23] are described by a 2 × 2 matrix pq∈{s,b},r∈{s,b} =(cid:18)pss psb pbs pbb(cid:19) (8) whose components are joint probabilities to observe the respective spiking (index "s") and/or bursting (index "b") states in x and y. The probabilities obey psb = pbs (due to stationarity), pb = pbb + psb, ps = pss + psb, thereby allowing to express all one- and two-time prob- abilities describing the dichotomous component in terms of two independent quantities, which for example can be a pair {ps, pss}, then psb = pbs = ps − pss, pbb = 1 − (pss + 2psb), (9a) (9b) or {pb, ρ} as in [19], where ρ is correlation coefficient defined by psb = pspb(1 − ρ). (10) In Section IV we justify the use of another effective pa- rameter ǫ (13) instead of ρ to determine time correlations in the dichotomous component. The two-time joint probabilities for the resulting pro- cess are then expressed as p(x 6= 1, y 6= 1) = psssxsy, p(x 6= 1, y = 1) = πsx, p(x = 1, y = 1) = p11, π = psss1 + psb, p(x = 1, y 6= 1) = πsy, p11 = psss2 1 + 2psbs1 + pbb. (11a) (11b) (11c) (11d) Note that the above notations can be applied to any subsystem instead of the whole system (with the same dichotomous component, as it is system-wide anyway). 3 hand and bursts on the other, because the latter are as- sumed to be correlated in time, unlike the former. That said, the "labeling" of bursts versus system-wide spikes exists in the model (by the state of the dichotomous com- ponent), but not in the realizations. Proceeding from the realizations, it must be possible to relabel a certain frac- tion of system-wide spikes into bursts (more precisely, into a time-uncorrelated portion thereof). Such relabel- ing would change both components of the model {V, S} (dichotomous and spiking processes), in particular dilut- ing the time correlations of bursts, without changing the actual realizations of the resultant process. This implies the existence of a transformation of model parameters which keeps realizations (i.e. the stochastic process as such) invariant. The derivation of this transformation is presented in Appendix A and leads to the following scaling sx = αs′ x, 1 − s1 = α(1 − s′ 1), ps′ = αps, ps′s′ = α2pss, (12a) (12b) (12c) (12d) where α is a positive scaling parameter, and all other probabilities are updated according to Eq. (9). The mentioned invariance in particular implies that any characteristic of the process must be invariant to the scaling (12a-d). This suggests a natural choice of a scaling-invariant effective parameter ǫ defined by pss = p2 s(1 + ǫ) (13) to determine time correlations in the dichotomous com- ponent. In conjunction with a second independent pa- rameter of the dichotomous process, for which a straight- forward choice is ps, and with full one-time probability table for spontaneous activity sx, these constitute a nat- ural full set of model parameters {sx, ps, ǫ}. The two-time probability table (8) can be expressed in terms of ps and ǫ by substituting Eq. (13) into Eq. (9): pq∈{s,b},r∈{s,b} =(cid:18) p2 s + ǫp2 pspb − ǫp2 s s (cid:19) . s pspb − ǫp2 p2 b + ǫp2 s (14) The requirement of non-negativeness of probabilities im- poses simultaneous constraints IV. MODEL PARAMETERS SCALING and ǫ ≥ −1 (15a) The spiking-bursting stochastic model as described in Section III is redundant in the following sense. In terms of the model definition, there are two distinct states of the model which equally lead to observing the same one-time state of the resultant process with 1's in all bits: firstly -- a burst, and secondly -- a system-wide simultaneous spike in the absence of a burst, which are indistinguish- able by one-time observations. Two-time observations reveal a difference between system-wide spikes on one ps ≤ ps max =( 1 1+ǫ(cid:16)1 −pǫ(cid:17) if − 1 ≤ ǫ < 0, if 1 1+ǫ ǫ ≥ 0, or, equivalently, − ǫ2 max ≤ ǫ ≤ ǫmax = pb ps . (15b) (16) Comparing the off-diagonal term psb in (14) to the def- inition of correlation coefficient ρ in (10), we get ǫ = ρ pb ps = ρ ǫmax, (17) thus the sign of ǫ has the same meaning as that of ρ. Hereinafter we limit ourselves to non-negative correla- tions ǫ ≥ 0. V. ANALYSIS OF THE EMPIRICAL "WHOLE MINUS SUM" MEASURE FOR THE SPIKING-BURSTING PROCESS In this Section we analyze the behavior of the "whole minus sum" empirical II [13] defined by Eqs. (3), (4) for the spiking-bursting model in dependence of the model parameters, particularly focusing on its transition from negative to positive values. Mutual information Ixy for two time instances x and y of the spiking-bursting process is expressed by insert- ing all one- and two-time probabilities of the process ac- cording to (6), (11) into the definition (1), (2). The full derivation is given in Appendix B and leads to an expres- sion which was used in [19] Ixy = 2(1 − s1){ps} + 2{p1} − (1 − s1)2{pss} − 2(1 − s1){π} − {p11}, where we denote for compactness {q} = −q log2 q. (18) (19) We exclude from further consideration the following degenerate cases which automatically give Ixy = 0 by definition (1): s1 = 1, or ps = 0, or ps = 1, or ρ = ǫ = 0, (20) where the former two correspond to a deterministic "al- ways 1" state for which all entropies in (1) are zero, and the latter two produce no predictability, which implies Hxy = Hx + Hy. The particular case s1 = 0 in (18) reduces to Ixys1=0 = 2(cid:0){ps} + {pb}(cid:1) −(cid:0){pss} + 2{psb} + {pbb}(cid:1) = I0(ps, ǫ), (21) which coincides with mutual information for the dichoto- mous component taken alone and can be seen as a func- tion denoted in (21) as I0(·,·) of just two independent parameters of the dichotomous component, for which we chose ps and ǫ as described in Section IV. Typical plots of I0(ps, ǫ) versus ps at fixed ǫ are shown with blue solid lines in Fig. 1. The general case (18) can be recovered back from (21) 1 = 0 by virtue of the scaling (12a-d), by assuming s′ 4 1 0.8 0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 1 FIG. 1. Blue solid lines -- plots of I0(ps, ǫ) versus ps varied from 0 to ps max as per (15), at ǫ = 0.01, 0.1, 0.2, 0.5, 1 (from right to left). Red dashed lines -- approximation (35). Red dots -- upper bounds of approximation applicability range (36c). in (12b) and substituting the corresponding scaled value ps′ = (1−s1)ps as per (12c) in place of the first argument of function I0(ps′ , ǫ) defined in (21), while parameter ǫ re- mains invariant to the scaling. This produces a simplified expression Ixy = I0(cid:0)(1 − s1)ps, ǫ(cid:1), (22) which is still exactly equivalent to (18). We emphasize that hereinafter expressions containing I0(·,·) like (22), (23), (30b) etc. imply that all probabilities in (21) must be expressed in terms of ps and ǫ, and ps in turn be accordingly substituted by the actual first argument of I0(·,·), e.g. by (1− s1)ps in (22). The same applies when the approximate expression for I0(·) (35) is used. Given a bipartition AB (see Section II), this result is applicable as well to any subsystem A (B), with s1 replaced by sA (sB) which denote the probability of a subsystem-wide simultaneous spike xA = 1 (xB = 1) in the absence of a burst, and with same parameters of the dichotomous component (here ps, ǫ). Then effective information (3) is expressed as Φeff = I0(cid:0)(1−s1)ps, ǫ(cid:1)−I0(cid:0)(1−sA)ps, ǫ(cid:1)−I0(cid:0)(1−sB)ps, ǫ(cid:1). (23) Hereafter in this section we assume the independence of spontaneous activity across the system, which implies sAsB = s1, Φeff = f (sA), then (23) turns into where f (s) = I0(cid:0)(1 − s1)ps, ǫ(cid:1) − I0(cid:0)(1 − s)ps, ǫ(cid:1) − I0(cid:0)(1 − s1/s)ps, ǫ(cid:1). (24) (25a) (25b) Note that the function I0(·,·) in (21) is defined only when the first argument is in the range (0, 1), thus the defini- tion domain of f (s) in (25b) is s1 < s < 1. (26) According to (4), the necessary and sufficient condition for the "whole minus sum" empirical II be positive is the requirement that Φeff be positive for any bipartition AB. Due to (25), this requirement can be written in the form min s∈{sA} f (s) > 0, (27) where {sA} is the set of sA values for all possible bipar- titions AB (if A is any non-empty subsystem, then sA is defined as the probability of spontaneous occurrence of 1's in all bits in A in the same instance of the discrete time). Expanding the set of s in (27) to the whole defini- tion domain of f (s) (26) leads to a sufficient (generally, stronger) condition for positive II min s1<s<1 f (s) > 0. (28) Note [24] that f (s) by definition (25b) satisfies f (s = s1) = f (s = 1) = 0, f ′(s = s1) > 0 and (due to the invariance to mutual renaming of subsystems A and B) f (s1/s) = f (s). The latter symmetry implies that the quantity of extrema on (s1, 1) must be odd, one of them always being at s = √s1. If the latter is the only ex- tremum, then it is a positive maximum, and (28) is thus fulfilled automatically. In case of three extrema, f (√s1) is a minimum, which can change sign. In both these cases the condition (28) is equivalent to the requirement f (√s1) > 0, which can be rewritten as g(s1) > 0, where (29) (30a) g(s1) = f (√s1) = I0(cid:0)(1 − s1)ps, ǫ(cid:1) − 2I0(cid:0)(1 − √s1)ps, ǫ(cid:1). (30b) The equivalence of (29) to (28) would be broken in case of 5 or more extrema. As suggested by numerical evi- dence [25], this exception never holds, although we did not prove this rigorously. Based on the reasoning above, in the following we assume the equivalence of (29) (and (30)) to (28). A typical scenario of transformations of f (s) with the change of s1 is shown in Fig. 2. Here the extremum f (√s1) (shown with a dot) transforms with the decrease of s1 from a positive maximum into a minimum, which in turn decreases from positive through zero into negative values. Note that by construction, the function g(s1) defined in (30b) expresses effective information Φeff from (3) for 5 0.006 0.004 0.002 0 −0.002 −0.004 −0.006 0 0.2 0.4 0.6 0.8 1 FIG. 2. Plots of f (s) on s1 < s < 1 with s1 = 0.03, 0.025, 0.0182919, 0.01 (from top to bottom) at ps = 0.7, ǫ = 0.1. For each value of s1, the extremum (√s1, f (√s1)) is indicated with a dot. a specific bipartition characterized by sA = sB = √s1. If such "symmetric" bipartition exists, then the value √s1 belongs to the set {sA} in (27), which implies that (29) (same as (30)) is equivalent not only to (28), but also to the necessary and sufficient condition (27). Otherwise, (28) (equivalently, (29) or (30)), formally being only suf- ficient, still may produce a good estimate of the neces- sary and sufficient condition in cases when {sA} contains values which are close to √s1 (corresponding to nearly symmetric partitions, if such exist). Except for the degenerate cases (20), g(s1) is negative at s1 = 0 g(s1 = 0) = −I0(ps, ǫ) < 0 and tends to +0 [26] with s1 → 1 − 0 as soon as (31) (32) lim s1→1−0 I0(cid:0)(1 − s1)ps, ǫ(cid:1) 2I0(cid:0)(1 − √s1)ps, ǫ(cid:1) = 2, hence g(s1) changes sign at least once on s1 ∈ (0, 1). Ac- cording to numerical evidence [27], we assume that g(s1) changes sign exactly once on (0, 1) without providing a rigorous proof for the latter statement (note however that for the asymptotic case (38) this statement is rigorous). In line with the above, the solution to (30a) has the form smin 1 (ps, ǫ) < s1 < 1, (33) 1 where smin (ps, ǫ) is the unique root of g(s1) on (0, 1). Several plots of smin (ps, ǫ) versus ps at ǫ fixed and versus ǫ at ps fixed, which are obtained by numerically solving for the zero of g(s1), are shown in Fig. 3 with blue solid lines. 1 Further insight into the dependence of mutual infor- mation Ixy (and, consequently, of Φeff and II) upon pa- rameters can be obtained by inserting the expressions 6 0.15 0.1 0.05 0 0 0.2 0.4 0.6 0.8 1 (a) (b) 0.05 0.04 0.03 0.02 0.01 0 0 0.2 0.4 0.6 0.8 1 FIG. 3. (a) Blue solid lines -- plots of smin (ps, ǫ) versus ps varied from 0 to ps max as per (15), at ǫ = 0.1, 0.5, 1 (from right to left). Red dashed line -- plot of the asymptotic formula (38). (b) Blue solid lines -- plots of smin (ps, ǫ) versus ǫ varied from 0 to ǫmax as per (16), at ps = 0.5, 0.6, 0.7 (from top to bottom). Vertical position of red dashed lines is the result of (38), horizontal span denotes the estimated applicability range (36b). 1 1 for the two-time probabilities (14) into the definition of I0(ps, ǫ) in (21) and expanding it in powers of ǫ (weak time correlation limit), which yields I0(ps, ǫ) = 1 2 log 2(cid:18) ps 1 − ps(cid:19)2 · ǫ2 + O(ǫ3). (34) Estimating the residual term (see details in Appendix C) indicates that the approximation by the leading term I0(ps, ǫ) ≈ is valid when ǫ2 2 log 2(cid:18) ps 1 − ps(cid:19)2 ǫ ≪ 1, ps(cid:19)2 ǫ ≪(cid:18) pb = ǫ2 max . (35) (36a) (36b) Solving (36b) for ps rewrites it in the form of an upper bound [28] for ps ps < . (36c) 1 1 +pǫ Note how inequalities (36b), (36c) compare to the formal upper bounds ǫmax in (16) and ps max in (15) which arise from the definition of ǫ (13) due to the requirement of positive probabilities. which do not depend on ǫ at all, this scaling of Φeff does not change the minimum information bipartition, finally implying that II also scales as ǫ2. That said, as factor ǫ2 does not affect the sign of Φeff, the lower bound smin in (33) exists and is determined only by ps in this limit. 1 Substituting the approximation (35) for I0(·) into the definition of g(s1) in (30b) after simplifications reduces the equation g(s1) = 0 to the following [29]: ps(√2 − 1) s1 − √s1 + (1 − ps)(√2 − 1) = 0 , (37) whose solution in terms of s1 on 0 < s1 < 1 equals smin , according to the reasoning behind Eq. (33). Solving 1 (37) as a quadratic equation in terms of √s1 produces a unique root on (0, 1), which yields smin 1 (ps)ǫ→0 =  1 −q1 − 4ps(1 − ps)(√2 − 1)2 2ps(√2 − 1) 2 .   (38) Result of (38) is plotted in Fig. 3 with red dashed lines: in panel (a) as a function of ps, and in panel (b) as hori- zontal lines whose vertical position is the result of (38), and horizontal span denotes the estimated applicability range (36b) (note that condition (36a) also applies, and becomes stronger than (36b) when ps < 1/2). Approximation (35) is plotted in Fig. 1 with red dashed lines along with corresponding upper bounds of approx- imation applicability range (36c) denoted by red dots (note that large ǫ violates (36a) anyway, thus in this case (36c) has no effect). Mutual information (35) scales with ǫ within range (36) as ǫ2 and vanishes with ǫ → 0. The same holds for effective information (23). Since the nor- malizing denominator in (4b) contains one-time entropies VI. COMPARISON OF INTEGRATED INFORMATION MEASURES In this Section we compare the outcome of two versions of empirical Integrated Information measures available in the literature, one being the "all-minus-sum" effec- tive information (3) from [13] which is used elsewhere in this study, and the other "decoder based" information as introduced in [16] and expressed by Eqs. (5a-c). We cal- culate both measures by their respective definitions using the one- and two-time probabilities from Eqs. (6a,b) and (11a-d) for the spiking-bursting model with N = 6 bits, assuming no spatial correlations among bits in spiking activity, with same spike probability P in each bit. In this case sx = P m(x)(1 − P )N −m(x), P = s 1 N 1 , (39) where m(x) is the number of ones in the binary word x. We consider only a symmetric bipartition with subsys- tems A and B consisting of N/2 = 3 bits each. Due to the assumed equal spike probabilities in all bits and in the absence of spatial correlations of spiking, this im- plies complete equivalence between the subsystems. In particular, in the notations of Sec. V we get sA = sB = √s1. s1 = sAsB, (40) This choice of the bipartition is firstly due to the fact that the sign of effective information for this bipartition determines the sign [30] of the resultant "whole minus sum" II. This has been established in Sec. V (see reason- ing behind Eqs. (27) -- (30) and further on); moreover, this effective information has been denoted in Eq. (30) as a function Φeff(AB) = g(s1), (41) which has been analyzed in Sec. V. Moreover, the choice of the symmetric bipartition is consistent with available comparative studies of II mea- sures [18], where it was substantiated by the conceptual requirement that highly asymmetric partitions should be excluded [2], and by the lack of a generally accepted spec- ification of minimum information bipartition; for further discussion, see [18]. We have studied the dependence of the mentioned ef- fective information measures upon spiking activity, which is controlled by s1, at different fixed values of the parame- ters ps and ǫ specifying the bursting component. Typical dependence of both measures upon s1, taken at ps = 0.6 with several values of ǫ, is shown in Fig. 4, panel (a). The behavior of the "whole minus sum" effective infor- mation Φeff (41) (blue lines in Fig. 4) is found to agree with the analytical findings of Sec. V: 1 • Φeff transitions from negative values to positive at a certain threshold value of s1 = smin , which is well approximated by the formula (38) when ǫ is small, as required by (36a,b); the result of Eq. (38) is indicated in each panel of Fig. 4 by an additional vertical grid line labeled smin on the abscissae axis, cf. Fig. 3; • Φeff reaches a maximum on the interval smin 1 and tends to zero (from above) at s1 → 1; • Φeff scales with ǫ as ǫ2, when (36a,b) hold. 1 < s1 < 1 7 To verify the scaling observation, we plot the scaled values of both information measures Φeff/ǫ2, Φ∗/ǫ2 in the panels (b) -- (d) of Fig. 4 for several fixed values of ps and ǫ. Expectedly, the scaling fails at ps = 0.7, ǫ = 0.4 in panel (d), as (36b) is not fulfilled in this case. Furthermore, the "decoder based" information Φ∗ (plotted with red lines in Fig. 4) behaves mostly the same way, apart from being always non-negative (which was one of key motivations for introducing this measure in [16]). At the same time, the sign transition point smin of the "whole minus sum" information associates with a rapid growth of the "decoder based" information. When s1 is increased towards 1, the two measures converge. Re- markably, the scaling as ǫ2 is found to be shared by both effective information measures. 1 VII. DISCUSSION In general, the spiking-bursting model is completely specified by the combination of a full single-time proba- bility table sx (consisting of 2N probabilities of all pos- sible outcomes, where N is the number of bits) for the time-uncorrelated spontaneous activity, along with two independent parameters (e.g. ps and ǫ) for the dichoto- mous component. This combination is, however, redun- dant in that it admits a one-parameter scaling (12) which leaves the resultant stochastic process invariant. Condition (30) was derived assuming that spiking ac- tivity in individual bits (i.e. nodes, or neurons) consti- tuting the system is independent among the bits, which implies that the probability table sx is fully determined by N spike probabilities for individual nodes. The con- dition is formulated in terms of ps, ǫ and a single pa- rameter s1 (system-wide spike probability) for the spon- taneous activity, thus ignoring the "internal structure" of the system, i.e. the spike probabilities for individual nodes. This condition provides that the "whole minus sum" effective information is positive for any bipartition, regardless of the mentioned internal structure. More- over, in the limit (36) of weak correlations in time, the inequality (30a) can be explicitly solved in terms of s1, producing the solution (33), (38). In this way, the inequality (33) together with the asymptotic estimate (38) supplemented by its applicabil- ity range (36) specifies the region in the parameter space of the system, where the "whole minus sum" II is posi- tive regardless of the internal system structure (sufficient condition). The internal structure (though still without spike correlations across the system) is taken into account by the necessary and sufficient condition (27) for positive II. The mentioned conditions were derived under the as- sumption of absent correlation between spontaneous ac- tivity in individual bits (24). If correlation exists and is positive, then s1 > sAsB, or sB < s1/sA. Then com- paring the expressions for Φeff (23) (general case) to (25) (space-uncorrelated case), and taking into account that (a) 0.1 0.08 0.06 0.04 0.02 0 0 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 (b) 0.1 0.2 0.3 0.4 8 0.2 0.15 0.1 0.05 0 0 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 0.1 0.2 0.3 0.4 0.1 0.2 0.3 0.4 (c) (d) FIG. 4. Comparison of two versions of empirical effective information for the symmetric bipartition -- "whole-minus-sum" measure (3) from [13] (blue lines) and "decoder based" information (5) from [16] (red lines) versus spiking activity parameter s1 at various fixed values of the bursting component parameters ps (indicated on top of the panels) and ǫ (indicated in the legends). Panel (a) -- unnormalized values, panels (b) -- (d) -- normalized by ǫ2. Threshold smin calculated according to (38) is shown in each panel with an additional vertical grid line. 1 I0(ps) is an increasing function, we find Φeff < f (sA), cf. (25a). This implies that any necessary condition for positive II remains as such. Likewise, in the case of neg- ative correlations we get Φeff > f (sA), implying that a sufficient condition remains as such. We found that II scales as ǫ2 for ǫ small (namely, within (36)) when other parameters (i.e. ps and spiking prob- ability table sx) are fixed. For the "whole minus sum" information, this is an analytical result. Note that the reasoning behind this result does not rely upon the as- sumption of spatial uncorrelation of spiking activity (be- tween bits) and thus applies to arbitrary spiking-bursting systems. According to a numerical calculation, this scal- ing is applicable to the "decoder based" information as well. Remarkably, II can not exceed the time delayed mutual information for the system as a whole, which in case of the spiking-bursting model in its present formulation is no greater than 1 bit. The present study substantiates, refines and quantifies 1 1 qualitative observations in regard to II in the spiking- bursting model which were initially made in [19]. The existence of lower bounds in spiking activity (character- ized by s1) which was noticed in [19] is now expressed in the form of an explicit inequality (33) with the estimate (38) for the bound smin . The observation of [19] that typically smin is mostly determined by burst probability and weakly depends upon time correlations of bursts also becomes supported by the quantitative result (33), (38). The model provides a basis for possible modifications in order to apply Integrated Information concepts to sys- tems exhibiting similar, but more complicated behavior (in particular, to neuron-astrocyte networks). Such mod- ifications might incorporate non-trivial spatial patterns in bursting, and causal interactions within and between the spiking and bursting subsystems. The model can also be of interest as a new discrete- state test bench for different formalizations of Integrated Information, while available comparative studies of II measures mainly focus on Gaussian autoregressive mod- els [17, 18]. ACKNOWLEDGMENTS This research was funded by the Ministry of Education and Science of the Russian Federation within Agreement No. 075-15-2019-871. Appendix A: Derivation of parameters scaling of the spiking-bursting model In order to formalize the reasoning in Section IV, we introduce an auxiliary 3-state process W with set of one- time states {s′, d, b}, where s′ and b are always inter- preted as spiking and bursting states in terms of Sec- tion III, and d is another state, which is assumed to produce all bits equal 1 like in a burst, but in a time- uncorrelated manner (which is formalized by Eq. (A4) below) like in a system-wide spike. When W is properly defined (by specifying all necessary probabilities, see be- low) and supplemented with a time-uncorrelated process S as a source of spontaneous activity for the state s′, these together constitute a completely defined stochastic model {W, S}. This 3-state based model may be mapped on equivalent (in terms of resultant realizations) 2-state based models as in Section III in an ambiguous way, because the state d may be equally interpreted either as a system-wide spike, or as a time-uncorrelated burst, thus producing two dif- ferent dichotomous processes (which we denote as V and V ′) for the equivalent spiking-bursting models. The rela- tionship between the states of W , V and V ′ is illustrated by the following diagram. V = s b (A1) As soon as d-states of W are interpreted in V as (spik- ing) s-states, the spontaneous activity process S accom- panying V has to be supplemented with system-wide spikes whenever W = d, in addition to the spontaneous activity process S′ for V ′. In order to maintain the ab- sence of time correlations in spontaneous activity (which is essential for the analysis in Section V), we assume time- uncorrelated choice between W = s′ and W = d when V = s (which manifests below in Eq. (A4)). Then the difference between the spontaneous components S and S′ comes down to a difference in the corresponding one-time probability tables sx and s′ x. In the following, we proceed from the dichotomous pro- cess V defined as in Section III, then define a consistent 3-state process W , and further obtain another dichoto- mous process V ′ for an equivalent model. Finally, we establish the relation between the corresponding proba- bility tables of spontaneous activity sx and s′ x. W = s′ d b V ′ = s′ z}{ {z} b′ 9 The first dichotomous process V has states denoted by {s, b} and is related to W according to the rule V = s when W = s′ or W = d, and V = b whenever W = b (see diagram (A1)). Assume fixed conditional probabilities p(W = s′ V = s) = α, p(W = d V = s) = β = 1 − α, which implies one-time probabilities for W as (A2a) (A2b) ps′ = αps, pd = βps. (A3) The mentioned requirement of time-uncorrelated choice between W = s′ and W = d when V = s is ex- pressed by factorized two-time conditional probabilities p(W = s′s′ V = ss) = α2, p(W = s′d V = ss) = αβ = p(W = ds′ V = ss), p(W = dd V = ss) = β2. (A4a) (A4b) (A4c) Given the two-time probability table for V (8) along with the conditional probabilities (A2), (A4), we arrive at a two-time probability table for W pq∈{s′,d,b},r∈{s′,d,b} = s′ d b s′ α2pss αβpss αpsb d αβpss β2pss βpsb pbb b αpbs βpbs . ! (A5) Note that (A5) is consistent both with (A3), which is ob- tained by summation along the rows of (A5), and with (8), which is obtained by summation within the line- separated cell groups in (A5): pss ≡ ps′s′ + ps′d + pds′ + pdd psb ≡ ps′b + pdb pbs ≡ pbs′ + pbd pbb ≡ pbb. (A6a) (A6b) (A6c) (A6d) Consider the other dichotomous process V ′ with states {s′, b′} obtained from W according to the rule V ′ = b′ when W = d or W = b, and V ′ = s′ whenever W = s′ (see diagram (A1)). The two-time probability table for V ′ is obtained by another partitioning of the table (A5) s′ d b pqr = s′ α2pss αβpss αpsb d αβpss β2pss βpsb b αpbs pbb βpbs ! (A7) with subsequent summation of cells within groups, which yields ps′s′ = α2pss, ps′b′ = α(βpss + psb) = pb′s′, pb′b′ = β2pss + 2βpsb + pbb. (A8a) (A8b) (A8c) The corresponding one-time probabilities for V ′ read ps′ = αps, pb′ = βps + pb. (A9a) (A9b) In order to establish the relation between the one-time probability tables of spontaneous activity sx and s′ x, we equate the resultant one-time probabilities of observing a given state x as per (6) for the two equivalent models {V, S} and {V ′, S′} p(x 6= 1) = pssx = ps′ s′ x, p(x = 1) = pss1 + pb = ps′ s′ 1 + pb′ . (A10a) (A10b) Taking into account (A9), we finally get sx = αs′ x, 1 − s1 = α(1 − s′ 1). (A11a) (A11b) Equations (A8), (A9) and (A11) fully describe the transformation of the spiking-bursting model which keeps the resultant stochastic process invariant by the construc- tion of the transform. Taking into account that the di- chotomous process is fully described by just two inde- pendent quantities, e.g. ps and pss, all other probabili- ties being expressed in terms of these due to normaliza- tion and stationarity, the full invariant transformation is uniquely identified by a combination of (A11a,b), (A8a) and (A9a), which together constitute the scaling (12). Note that parameter α within its initial meaning (A2) may take on values in the range 0 < α ≤ 1 (case α = 1 producing the identical transform). That said, in terms 10 of the scaling (12a-d), all values α > 0 are equally possible, so that mutually inverse values α = α1 and α = α2 = 1/α1 produce mutually inverse transforms. Appendix B: Expressing mutual information for the spiking-bursting process One-time entropy Hx for the spiking-bursting process is expressed by (2) with probabilities p(x) taken from (6): Hx =Xx {p(x)} =Xx {pssx} + {p1} − {pss1}, (B1) where the additional terms besides the sum over x ac- count for the specific expression (6b) for p(x = 1). Using the relation {ab} ≡ a{b} + {a}b, (B2) which is derived directly from (19), and collecting similar terms, we arrive at Hx = psHs − ps{s1} + (1 − s1){ps} + {p1}, (B3) where Hs is the entropy of the spiking component taken alone Hs =Xx {sx}. (B4) Two-time entropy is expressed similarly, by substitut- ing probabilities p(xy) from (11) into the definition of entropy and taking into account the special cases with x = 1 and/or y = 1: Hxy =Xxy {p(xy)} =Xxy {psssxsy} −Xx −Xy {psssxs1} +Xx {psss1sy} +Xy {πsx} {πsy} + {psss2 1} − 2{πs1} + {p11}. (B5) Further, applying (B2) and using the notation (B4), we find sxsy = pss · 2Hs + {pss}, {sxsy} + {pss}Xxy Xxy {psssxsy} = pssXxy where we used the reasoning thatPxy{sxsy} is the two- time entropy of the spiking component taken alone, which is (due to the postulated absence of time correlations in it) twice the one-time entropy Hs (this of course can equally be found by direct calculation). Similarly, we get (B6a) also and exactly the same expression for Py{psss1sy}, and Xy {sx} + {π}Xx {πsx} = πXx {πsy} =Xx (B6c) sx = πHs + {π}. Substituting (B6a-c) into (B5), using (B2) where ap- plicable, and collecting similar terms with the relation pss + π − psss1 ≡ ps taken into account, we arrive at Xx {psssxs1} = psss1Xx {sx} + {psss1}Xx = psss1Hs + {psss1} sx (B6b) Hxy = 2psHs + (1 − s1)2{pss} − 2ps{s1} + 2(1 − s1){π} + {p11}. (B7) (B8) Finally, the expression (18) for mutual information is obtained by inserting (B3) and (B8) into the definition (1), with stationarity Hy = Hx taken into account. and in the case ξ < 0 to Appendix C: Expanding I0 in powers of ǫ where Taylor series expansion for a function f (x) up to the quadratic term reads 1 3 Φ(cid:18)ξ x0(cid:19) ≪ 1, Φ(ζ) = ζ (1 − ζ)2 . 11 (C8a) (C8b) f (x0 + ξ) = f (x0) + f ′(x0)ξ + f ′′(x0) ξ2 2 + R(ξ). (C1) Replacing Φ(·) in (C8a) by its linearization Φ(ζ) ≈ ζ for small ζ, we reduce both (C7) and (C8a) to a single condition The remainder term R(ξ) can be represented in the La- grange's form as ξ ≪ 3x0. (C9) We use these considerations to expand in powers of ǫ the function I0(ps, ǫ) defined in (21) with pss, psb, pbb substituted by their expressions in terms of ǫ according to (14). We note that the braces notation {·} defined in (19) is expressed via the function f (x) from (C4) as (C3a) Expanding this way the subexpressions of (21) R(ξ) = f ′′′(c) ξ3 6 , (C2) where c is an unknown real quantity between x0 and x0 + ξ. The function f (x) can be approximated by omit- ting R(ξ) in (C1) if R(ξ) is negligible compared to the quadratic term, for which it is sufficient that for any c between x0 and x0 + ξ, namely for f ′′(x0) ξ2 2 (cid:12)(cid:12)(cid:12)(cid:12) ξ3 f ′′′(c) (cid:12)(cid:12)(cid:12)(cid:12) 6 (cid:12)(cid:12)(cid:12)(cid:12) ≪(cid:12)(cid:12)(cid:12)(cid:12) c ∈((x0, x0 + ξ), (x0 − ξ, x0), if ξ > 0, if ξ < 0. (C3b) Consider the specific case f (x) = −x log x, x > 0, (C4) for which we get , 1 x f ′′′(x) = f ′′(x) = − 1 f ′(x) = − log x − 1, x2 . (C5) As long as f ′′′(x) is a falling function for any x > 0, fulfilling (C3a) at the left boundary of (C3b) (at c = x0 if ξ > 0, and at c = x0 − ξ if ξ < 0) makes sure (C3a) is fulfilled in the whole interval (C3b). Precisely, the requirement is 1 x2 0 (cid:12)(cid:12)(cid:12)(cid:12) 1 (x0 − ξ)2 (cid:12)(cid:12)(cid:12)(cid:12) 1 x0 1 x0 ξ3 ξ3 6 (cid:12)(cid:12)(cid:12)(cid:12) ≪(cid:12)(cid:12)(cid:12)(cid:12) 6 (cid:12)(cid:12)(cid:12)(cid:12) ≪(cid:12)(cid:12)(cid:12)(cid:12) ξ2 ξ2 2 (cid:12)(cid:12)(cid:12)(cid:12) 2 (cid:12)(cid:12)(cid:12)(cid:12) which in the case ξ > 0 reduces to , , if ξ > 0, (C6a) if ξ < 0, (C6b) ξ 3x0 ≪ 1, (C7) {q} = f (q) log 2 . s + ǫp2 {pss} = {p2 s}, {psb} = {pspb − ǫp2 s}, b + ǫp2 {pbb} = {p2 s}, (C10) (C11a) (C11b) (C11c) we find by immediate calculation that the zero-order and linear in ǫ terms vanish, and the quadratic term yields (35). The condition (C9) has to be applied to all three subexpressions (C11a-c). Omitting the insignificant fac- tor 3 in (C9), we obtain the applicability conditions ǫp2 ǫp2 ǫp2 s ≪ p2 s, s ≪ pspb, s ≪ p2 b, which is equivalent to ǫ ≪ 1, pb ǫ ≪ ps ǫ ≪ ǫ2 max, = ǫmax, (C12a) (C12b) (C12c) (C13a) (C13b) (C13c) where the notation ǫmax from (16) is used. We note that when ǫmax < 1, the condition (C13c) is the strongest among (C13a-c); when ǫmax > 1, the condition (C13a) is the strongest. Therefore, in both cases (C13b) can be dropped, thus producing (36). [1] G. Tononi, BMC Neurosci. 5, 42 (2004). [2] D. Balduzzi and G. Tononi, PLoS Comput. Biol. 4, [17] M. Tegmark, PLoS Comput. Biol. 12, e1005123 (2016). [18] P. Mediano, A. Seth, and A. Barrett, Entropy 21, 17 e1000091 (2008). (2019). [3] G. Tononi, Archives italiennes de biologie 150, 293 [19] O. Kanakov, S. Gordleeva, A. Ermolaeva, S. Jalan, and 12 (2012). [4] M. Oizumi, L. Albantakis, and G. Tononi, PLoS Com- put. Biol. 10, e1003588 (2014). [5] G. Tononi, Biol. Bull. 215, 216 (2008). [6] A. Peressini, J. Conscious. Stud. 20, 180 (2013). [7] N. Tsuchiya, S. Taguchi, and H. Saigo, Neurosci. Res. 107, 1 (2016). [8] G. Tononi, M. Boly, M. Massimini, and C. Koch, Nat. Rev. Neurosci. 17, 450 (2016). [9] R. Norman and A. Tamulis, J. Comput. Theor. Nanosci. 14, 2255 (2017). [10] D. Engel and T. W. Malone, PLoS ONE 13, e0205335 (2018). [11] P. A. M. Mediano, J. C. Farah, arXiv han, arXiv:1606.08313 [q-bio.NC]. e-prints , arXiv:1606.08313 and M. Shana- (2016), [12] D. Toker and F. T. Sommer, PLoS Comput. Biol. 15, e1006807 (2019). [13] A. B. Barrett and A. K. Seth, PLoS Comput. Biol. 7, e1001052 (2011). [14] V. Griffith, arXiv e-prints , arXiv:1401.0978 (2014), arXiv:1401.0978 [cs.IT]. [15] M. Oizumi, N. Tsuchiya, and S.-i. Amari, Proc. Natl. Acad. Sci. USA 113, 14817 (2016). [16] M. Oizumi, S.-i. Amari, T. Yanagawa, N. Fujii, and N. Tsuchiya, PLoS Comput. Biol. 12, e1004654 (2016). A. Zaikin, Phys. Rev. E 99, 012418 (2019). [20] A. Araque, G. Carmignoto, P. G. Haydon, S. H. Oliet, R. Robitaille, and A. Volterra, Neuron 81, 728 (2014). [21] A. B. Barrett, Phys. Rev. E 91, 052802 (2015). [22] In a real network "simultaneous" implies occuring within the same time discretization interval [19]. [23] In a neural network these correlations are conditioned by burst duration [19]; e.g., if this (in general, random) du- ration mostly exceeds τ , then the correlation is positive. [24] All mentioned properties and subsequent reasoning can be observed in Fig. 2, which shows a few sample plots of f (s). [25] The equivalence of (28) to (30) was confirmed up to ma- chine precision for each combination of ps ∈ [0.01, 0.99] and ρ ∈ [0.01, 1] (both with step 0.01). [26] Notations −0 and +0 denote the left ang right one-sided limits. [27] This statement was confirmed up to machine precision for each combination of ps ∈ [0.01, 0.99] and ρ ∈ [0.01, 1] (both with step 0.01). [28] The use of '≪' sign is not appropriate in (36c), because this inequality does not imply a small ratio between its left-hand and right-hand parts. [29] See comment below Eq. (22). [30] Although the actual value of II is determined by the min- imal information bipartition which may be different.
1502.07907
1
1502
2015-02-27T14:22:14
Closed Form Jitter Analysis of Neuronal Spike Trains
[ "q-bio.NC" ]
Interval jitter and spike resampling methods are used to analyze the time scale on which temporal correlations occur. They allow the computation of jitter corrected cross correlograms and the performance of an associated statistically robust hypothesis test to decide whether observed correlations at a given time scale are significant. Currently used Monte Carlo methods approximate the probability distribution of coincidences. They require generating $N_{\rm MC}$ simulated spike trains of length $T$ and calculating their correlation with another spike train up to lag $\tau_{\max}$. This is computationally costly $O(N_{\rm MC} \times T \times \tau_{\max})$ and it introduces errors in estimating the $p$ value. Instead, we propose to compute the distribution in closed form, with a complexity of $O(C_{\max} \log(C_{\max}) \tau_{\max})$, where $C_{\max}$ is the maximum possible number of coincidences. All results are then exact rather than approximate, and as a consequence, the $p$-values obtained are the theoretically best possible for the available data and test statistic. In addition, simulations with realistic parameters predict a speed increase over Monte Carlo methods of two orders of magnitude for hypothesis testing, and four orders of magnitude for computing the full jitter-corrected cross correlogram.
q-bio.NC
q-bio
Closed Form Jitter Analysis of Neuronal Spike Trains Daniel M. Jeck1,2 and Ernst Niebur1,3 1Zanvyl Krieger Mind/Brain Institute, Johns Hopkins University, Baltimore, MD 2Department of Biomedical Engineering, Johns Hopkins 3Solomon Snyder Department of Neuroscience, Johns Hopkins University, Baltimore, MD University, Baltimore, MD July 16, 2018 5 1 0 2 b e F 7 2 ] . C N o i b - q [ 1 v 7 0 9 7 0 . 2 0 5 1 : v i X r a 1 Closed Form Jitter Analysis Running Title Closed Form Jitter Analysis Address correspondence to Daniel Jeck The Johns Hopkins University Krieger Mind/Brain Institute 3400 N. Charles Street Baltimore, MD 21218 (email: [email protected]) Keywords Jitter, Spike timing, Synchrony, Cross Correlation Abstract Interval jitter and spike resampling methods are used to analyze the time scale on which temporal correlations occur. They allow the computation of jitter corrected cross correlograms and the per- formance of an associated statistically robust hypothesis test to de- cide whether observed correlations at a given time scale are signifi- cant. Currently used Monte Carlo methods approximate the proba- bility distribution of coincidences. They require generating NMC sim- ulated spike trains of length T and calculating their correlation with another spike train up to lag τmax. This is computationally costly O(NMC × T × τmax) and it introduces errors in estimating the p value. Instead, we propose to compute the distribution in closed form, with a complexity of O(Cmax log(Cmax)τmax), where Cmax is the maximum possible number of coincidences. All results are then exact rather than approximate, and as a consequence, the p-values obtained are the theoretically best possible for the available data and test statis- tic. In addition, simulations with realistic parameters predict a speed increase over Monte Carlo methods of two orders of magnitude for hypothesis testing, and four orders of magnitude for computing the full jitter-corrected cross correlogram. 2 Closed Form Jitter Analysis 1 Introduction Though there are many means by which neurons communicate, using both chemical and electrical mechanisms, most attention has been paid to series of action potentials (spikes). It is known that in some cases the detailed time structure of these spike trains is used for information transmission while in others, only the overall number of spikes in some interval seems to be important but not their position in the interval. Examples of the first type are various kinds of "temporal coding" schemes proposed for different neural system and for different functional roles, e.g. refs Abeles (1991); Singer and Gray (1995); Riehle et al. (1997); Niebur et al. (1993); Softky (1995); Steinmetz et al. (2000), while the latter is the well-known rate-code mechanism, e.g. refs. Adrian and Zotterman (1926); Shadlen and Newsome (1998). To distinguish between these two possibilities, it is necessary to find whether reproducible correlations at the relevant time scale are present in neuronal data. One common way to approach this problem is to use auto- or cross-correlation functions as test statistics. Then one can (a) search for non-trivial structure in the function, like deviations from uniformity, and (b) detect whether there are differences between these functions in different experimental (e.g. behavioral) conditions. The situation is complicated by the influence of rate variations on the raw correlations. To increase the signal-to-noise ratio, correlations are typically computed as averages over many trials. Changes in the behavioral state of an animal, e.g. due to onset of sensory stimuli or motor responses that occur always at the same time during a trial, typically result in changes in neural firing rates which are common to many neurons. While these are genuine correlations, they are unrelated to the neuronal coding question. Different techniques have been developed to remove them, e.g. subtraction of a "shuffle predictor" Perkel et al. (1967), the average of cross correlations between spike trains from permuted trials1. While this correction removes correlations that are time-locked to trial onset, it was later pointed out that peaks in the correlation function that may be taken as indicative of correlated firing (e.g. at zero lag) can also be caused by slow rate covariations Brody (1998, 1999). After finding a significant peak in the cross correlation function, this ambiguity can be addressed by analyzing the time scale at which the 1A "shift predictor" is very similar but the correlation function is computed from trials that immediately follow each other, rather than from randomly selected trials. 3 Closed Form Jitter Analysis measured correlation arises. It was pointed out more recently Amarasingham et al. (2012) that the null hypothesis of spike trains being independent in earlier work Perkel et al. (1967) is useless if their mean rates co-vary. Rate co-variation of means the spike trains are not independent which leads to its immediate rejection of the null without providing any further insight. Amarasingham and his colleagues instead proposed a more detailed null hypothesis, namely that within an in- terval of width ∆, the exact location of spikes does not matter. Then, under the null hypothesis, simulated spike trains can be generated by modifying the spike times of an original measured spike train within a range of ∆. The cross correlations obtained from these modified ("jittered") spike trains are then compared to those obtained from the original. If significant differences are found, the null hypothesis is rejected and it is likely that non-random correlations at time scales ≤ ∆ are present in the data. Additionally, this method gives rise to the computation of jitter-corrected cross correlograms, which have been used to compare changes in synchrony across experimen- tal conditions (Hirabayashi et al., 2013a,b; Smith et al., 2013). Because the method relies on repeated simulation of spike trains, it will be referred to as the Monte Carlo jitter method for the purposes of this paper. While the Monte Carlo jitter method (described fully in Section 2.1) is useful and easily generalized to complex statistical tests and hypotheses, its practical utility is limited by the inherent trade-off between accuracy and computation time in all Monte Carlo methods. As we will show in Section 3.2, the computation time may be prohibitively long, and even at this cost, it will only be a numerical approximation of the true solution. In the case where the test statistic is the cross-correllation value at a single lag, the p value can be computed exactly, as was shown by Harrison (2013). In the present study, we therefore explore the benefits of computing in closed form the distribution which is only approximated by the Monte Carlo simulations. Accordingly we refer to this method, described in Section 2.2, as the closed form jitter method. In addition to computing the p value for rejecting the null hypothesis exactly we show that the computation of the jitter-corrected cross correlogram follows readily from that derivation. The computational performance of the closed form jitter and Monte Carlo jitter methods are compared theoretically (as computational complexity) in Section 3.1 and practically (as computational time) in Section 3.2. 4 Closed Form Jitter Analysis Length of binned spike trains Two spike trains Correlation Lag Correlation of X and Y Jitter interval width Index over Monte Carlo simulated signals Index over jitter intervals Number of spikes in X in interval j ith Monte Carlo simulated signal Number of Monte Carlo simulated signals Correlation of X MC Number of cases where C MC p Value for correlation at lag τ T X, Y τ C(τ ) ∆ i j N(X, j) X MC NMC C MC i Rτ pτ JCCG(τ ) Jitter-corrected Cross Correlogram Pτ (C MC) Monte Carlo estimate of the distribution of correlations at lag τ C int Pτ (C(τ )) True distribution of correlations at lag τ τmax Nmax Cmax Maximum τ value processed Maximum value of N(X, j) or N(Y, j) Maximum possible number of coincidences Number of coincidences in the jth interval (τ ) > C(τ ) i and Y i i j Table 1: Glossary. Variables are listed in the order in which they are introduced. 5 Closed Form Jitter Analysis 2 Materials and Methods 2.1 The Monte Carlo Jitter Method Utilizing the Monte Carlo jitter method (Amarasingham et al., 2012), it is possible to determine whether correlations arise from fine temporal structure or larger scale variations, sometimes referred to as rate covariations. This determination is made by comparing a test statistic (in this case cross cor- relations) of an original pair of spike trains against those computed from a set of jittered spike trains as described below. The jitter method, like cross correlation, operates on binned spike trains which we take as binary signals with values 0 and 1 and integer arguments 0 to T − 1, where T is the number of bins in the spike train. The binary assumption implies that the bin size is small enough (typically 1ms or so) such that two spikes cannot be recorded in a single time bin. A sufficiently small bin size can always be chosen since there are limits on the minimal inter-spike interval time due to the refractory period of the neurons in question. Let X(t) and Y (t) be two such binned spike trains. The processing then consists of the following steps: 1. Compute the cross correlation C(τ ) between the original X and Y , C(τ ) =Xt X(t − τ )Y (t) where the sum runs from 0 to T − 1 and X is assumed to be 0 if its argument is outside that range. 2. Subdivide one of the signals, say X, into intervals of width ∆. 3. Count the number of spikes in each interval of X. In interval j this is, N(X, j) = (j+1)∆−1 Xk=j∆ X(k) (1) 4. For X generate NMC Monte Carlo simulated signals {X MC }, in which the spike counts for each interval are the same as in the corresponding interval in X, such that N(X MC , j) = N(X, j) for all i, j. However, now spike times within the interval are all equally likely. Spike times should be sampled without replacement to ensure that the spike count stays constant without putting multiple spikes in a single bin. i i 6 Closed Form Jitter Analysis 5. Compute the cross correlation C MC i (τ ) for lag time τ between each and the second spike train Y to get an estimate of the distribution X MC Pτ (C MC) of cross correlation values for each time lag τ . i C MC i (τ ) =Xt X MC i (t − τ )Y (t) 6. Let Rτ be the number of simulations where C MC i (τ ) ≥ C(τ ). Then the p value for a given lag τ is computed as pτ = Rτ + 1 NMC + 1 7. If desired, a jitter-corrected cross correlogram (JCCG), defined using the expectation operator E[·], can be computed as JCCG(τ ) = C(τ ) − E[CMC(τ )] ≈ C(τ ) − 1 NMC Xi CMC i (τ ) (2) where the approximation approaches equality with NM C → ∞. Jittering the spikes within an interval of size ∆ destroys all correlations at time scales within this interval. The cross correlations computed from the jittered spike trains therefore are not correlated on time scales ∆ or smaller, and Pτ (C MC) is the distribution of correlations at time lag τ obtained under the null hypothesis that correlations at time scales ≤ ∆ are indistinguishable from random correlations. If the measured cross correlation C(τ ) is signif- icantly outside this distribution, we have to reject the null hypothesis and we conclude that nonrandom correlations at lag τ with time scales ≤ ∆ are found in the observed spike trains. If, on the other hand, the observed cor- relation is consistent with what is seen in the distribution of jittered spike trains, then we cannot reject the null hypothesis. This means we cannot exclude that the observed synchrony is caused by correlations on time scales outside the jittered range, in other words that the observed correlation at lag time τ is caused by rate variations on time scales greater than ∆. In practice, X(t) and Y (t) do not have to be gathered from a continuous block of time. In the case of multiple trials of the same experimental con- dition, it may be useful to concatenate the recorded spike trains (possibly 7 Closed Form Jitter Analysis after removing sections of them, like those recorded during stimulus onsets). In doing so, a period of no spiking of width τmax (the largest correlation lag of interest) should be added between the trials so that correlations between trials don't affect the outcome of the jitter procedure. As mentioned in Section 1, the practical utility of the Monte Carlo method is limited by the trade-off between accuracy and computation time inherent in all Monte Carlo algorithms. Furthermore, in practice a single set of Monte Carlo simulations is often generated for many hypothesis tests (i.e. tests at multiple lags), introducing potential dependencies between the different tests when they should be treated independently 2. In order to avoid both of these issues, the probability distribution Pτ (C MC) can be computed exactly and independently for each time lag as described in the following. 2.2 Closed Form Computation 2.2.1 Probability Distribution For One Interval First, let us consider a single interval consisting of ∆ time bins. For example, if spike times have been binned to 2 ms, for an interval of width 20 ms we have ∆ = 10. Since time has been discretized, it is still possible to discuss this unitless value as a length of time, a time scale, or a interval width for a given bin size. As before, we assume that the sequence is binary, so each bin has either zero or one spike. This is true even when the spike times are jittered because spike times are sampled without replacement. For this single interval, the probability of a given number of coincidences occurring is determined by three values: ∆, N(X, j) the number of spikes in interval j of spike train X, and N(Y, j) the number of spikes in interval j of spike train Y. As a first step, we count the number of perfect coincidences, in which one spike occurs in both X and Y within the same time bin, meaning τ = 0. Using the standard notation of(cid:0)a b), we find that there are (cid:0) ∆ each of them falls into one of these empty bins is (cid:0)∆−N (Y,j) b(cid:1) for the combinatorial operation (a choose N (X,j)(cid:1) ways to distribute N(X, j) spikes in ∆ N (X,j) (cid:1). These are all available bins. The number of empty (spike-less) bins in spike train Y is [∆ − N(Y, j)]. The number of ways to distribute N(X, j) spikes such that 2The procedure of generating one set of spike trains for multiple lags is appropriate inappropriate only if each lag is being tested independently. If the test statistic is the sum of C(τ ) over a range of lag values, a single set of simulated spike trains is appropriate.f 8 Closed Form Jitter Analysis possible cases in which a coincidence is avoided. The probability that zero coincidences occur in the j-th interval is therefore P (C int j = 0∆, N(X, j), N(Y, j)) = (cid:0)∆−N (Y,j) N (X,j) (cid:1) (cid:0) ∆ N (X,j)(cid:1) where C int j is the number of coincidences in this interval. (3) We can generalize equation 3 to a non-zero number c of coincidences by breaking the numerator up into the number of ways that c spikes can coincide with the spikes in Y , and N(X, j) − c spikes coincide with the gaps (or non-spikes) in Y . We can thus compute a probability distribution for each interval j, P (C int j = c∆, N(X, j), N(Y, j)) = (cid:0)∆−N (Y,j) N (X,j)−c(cid:1)(cid:0)N (Y,j) (cid:1) (cid:0) ∆ N (X,j)(cid:1) c (4) where we follow the customary convention of setting the value of a "choose" operation to zero if either of its arguments is negative, or if its upper argu- ment is less than the lower. If this happens in the numerator of eq. 4, the probability on the left hand side becomes zero. Of course, the denominator is always positive since N(X, j) ≤ ∆. This is a hypergeometric distribution. Equation 4 is easily generalized to nonzero values of τ by applying the analysis leading to it to a shifted version of Y . For the computation of N(Y, j), this implies adding τ to the summation limits in eq. 1. As with other correlation algorithms, the boundaries of finite spike trains (beginning and end) result in fewer intervals to analyze as τ gets further away from zero. Thus generalizing eq. 4 to non-zero τ , we denote the resulting number j (τ ) and the associated probability distributions as P int of coincidences as C int . τ 2.2.2 Jitter-Corrected Cross Correlation Once we have the analytical probability distribution for the correlations, we can obtain all relevant quantities to characterize the pairwise correlations be- tween two spike trains. It is straightforward to compute the commonly used jitter-corrected cross correlogram (e.g., Hirabayashi et al., 2013a,b; Smith et al., 2013) which shows the correlation function after all correlations on time scales longer than ∆ have been removed. It is defined in analogy to equation 2 where the expectation value of the stochastic solution, E[C MC(τ )], was used. 9 Closed Form Jitter Analysis We can replace this approximation by the exact solution E[C int(τ )]. Fur- thermore, by the null hypothesis each interval is conditionally independent based on the spike counts. Therefore, the JCCG can be computed without approximation by JCCG(τ ) = C(τ ) − E"Xj Cint j (τ )# = C(τ ) −Xj E(cid:2)Cint j (τ )(cid:3) (5) which, as should be remembered, is computed for a specific jitter interval width ∆. The expectation on the right can either be calculated for each window as N(x, j) × N(y, j)/∆. The jitter-corrected cross correlogram is used, for instance, when the scientific question of interest is whether there are significant changes in syn- chrony between conditions, rather than a test of the presence or absence of synchrony. It is then used as part of a bootstrap statistical test in which the observed pairwise correlation is compared with the distribution obtained from eq. 5. 2.2.3 Probability Distribution For Spike Train τ One can also obtain the probability distribution for the entire signal Pτ (C(τ )) as the convolution of the individual probability distributions for all intervals, P int . This is identical to computing Pτ (C MC) from Section 2.1 with an infinite number of Monte Carlo simulations for each value of τ . One can then evaluate how likely it is that the observed cross correlation C(τ ) is explained by this probability distribution. The likelihood p that this is the case is obtained as the integral of the probability density function exceeding C(τ ), as in pτ = ∞ Xc=C(τ ) Pτ (c) (6) 3 Results 3.1 Computational Complexity In many situations, the statistical distributions underlying the phenomena under study are complicated or unknown and performing Monte Carlo simu- lations are the only way to make progress, even though it may be costly and 10 Closed Form Jitter Analysis it introduces additional randomness in the processing. In the case considered here (binary spike trains, null hypothesis of uniform spike time distribution in fixed interval, cross correlation test statistic), the distribution Pτ (C) can be computed directly, using the closed form jitter method described above, without the need for repeated simulations. This section will compare the computational complexity of using the Monte Carlo jitter method against the direct computation using the closed form jitter method. 3.1.1 Monte Carlo Method In the Monte Carlo algorithm, the data generation step requires a permuta- tion of ∆ data points for each interval and simulation. Since a single per- mutation operation has a computational complexity O(∆), and ∆ times the number of intervals is the length of the signal T , generating the set of Monte Carlo simulations {X MC } is O(NMC ×T ). The complexity of cross correlation or convolution of two signals with lengths T is O(T × log T ), assuming an FFT-based method (Cooley and Tukey, 1965) is used. So computing the full Monte Carlo probability distribution for all values of τ is O(NMC×T ×log T ). In many cases, not all values of τ are needed. If the correlation is computed only for the subset of delay values from 0 to τmax, the complexity for the Monte Carlo jitter method is i O(NMC × T × τmax) Computing the jitter-corrected cross correlogram by this method only requires one additional sum, with complexity O(NMC × τmax) so the total complexity remains unchanged. 3.1.2 Closed Form Probability Distribution To compute the exact probability distribution with the closed form jitter method, note that the values of the distribution can be precomputed based on the maximum values of N(X, j) and N(Y, j) over all j; call this maxi- mum Nmax. Also, n choose k operations can be as fast as O(min(k, n − k)) (Manolopoulos, 2002). Therefore, a three dimensional table of all possible values of P (C int N(X, j), N(Y, j)) can be precomputed and then looked up for each interval. Generating this table requires up to Nmax different values of C, Nmax values of N(X, j), and Nmax values of N(Y, j). Computing each value requires three n choose k operations, which are on the order of O(Nmax), j 11 Closed Form Jitter Analysis 4). While the ex- so the total computation of the probability table is O(Nmax ponent is high, the expression does not have any dependence on the length of the signal and, furthermore, Nmax ≤ ∆ is a small number in essentially all cases of interest. In practice, for analyzing neurophysiological data it is rare that a time resolution finer than 1 ms is needed, or controlling for cross correlations at time scales larger than approximately 100 ms (i.e. ∆ ≈ 100). The full lookup table is therefore maximally a 100 × 100 × 100 matrix, which requires negligible resources to compute and store. To compute the combined probability distribution Pτ (C) over all inter- vals, all interval probability distributions Pτ (C int j ) must be convolved, and the computational complexity of the problem is dominated by these convo- lution operations. As will be shown, we can improve performance by taking advantage of the structure of the problem at hand, since many of the convo- lution operations are identical. As a result, the convolutions can be grouped together based on N(X, j) and N(Y, j) and quickly combined so that O(T ) 2) convolutions. This can be done by the convolutions will turn into O(Nmax following procedure: 1. Take the Fast Fourier Transform (FFT) of Pτ (C int j for each encountered value of N(X, j) and N(Y, j). N(X, j), N(Y, j)) 2. Raise each complex frequency spectrum value to a power equal to the number of times that the (N(X, j), N(Y, j)) pair appears. 3. Multiply these frequency spectra. 4. Take the inverse FFT of the result to get the final probability distribu- tion Pτ (C) and compute pτ as in equation 6. 5. Repeat steps 2 through 4 for each value of τ to be tested. The FFT operations in step 1 must be zero-padded up to the maximum number of coincident spikes Cmax to account for the highest possible number of synchronous spikes in the combined probability distribution. Therefore the FFT operation in step 1 is O(Cmax × log(Cmax)). In step 2, exponentiation is 2) exponents to be taken, repeated τmax O(1). However there are O(T × Nmax 2) multiplications, again repeated times in step 5. Step 3 requires O(T ×Nmax τmax times. In step 4, the length of the spectral signal (to be inverted by FFT) is Cmax, so the operation is O(Cmax × log(Cmax)) repeated τmax times. When 12 Closed Form Jitter Analysis combining these steps, note that Cmax is proportional to T . However Cmax will be used when relevant because it captures the frequency dependence of the computation time. Therefore the total computational complexity is O(Cmax × log(Cmax) × τmax) . Note that because the zero-frequency component of a probability distribu- tion is always exactly unity, the inverse FFT computation will have accuracy limited by the precision of the numerical system. In practice this implies that p values less than 10−13 will not be estimated accurately. 3.1.3 Jitter-Corrected Cross Correlation Both the complexity analysis and the actual computation of the jitter-corrected cross correlogram is much simpler than that of the probability distribution. We generate a lookup table of possible E(cid:2)C int(cid:3) values and, from equation 5, the jitter-corrected correlogram can be computed at a speed of O(T × τmax) . 3.2 Computational Execution Time For practical applications, consumption of resources is an important limita- tion for any computational method. For the size of problems encountered in typical neurophysiological experiments, the only limiting resource is execu- tion time. To compare the performance of the Monte Carlo jitter and the closed form method, the two algorithms were run side by side in the MAT- LAB environment (Mathworks, Natick MA). Synthetic spike trains were gen- erated that varied in both frequency of spiking (5 to 500 Hz) and length (1 to 91 seconds). For each (time, frequency) condition, 50 spike trains were generated, binned to 1 ms, and the average processing time was computed. Processing was performed with τmax = 100ms and ∆ = 20. All computations were performed on an Intel i7 920 processor with 12 GB of RAM running Linux Ubuntu 12.04. For the Monte-Carlo method, NMC was set to 1000. This selection of NMC is unrealistically low for two reasons. First, it can at best result in a 13 Closed Form Jitter Analysis Bonferroni corrected p value of 0.201 due to the 201 p values being tested in the range of −τmax to τmax. As the execution for NMC = 1000 already takes 5.7 days to run, increasing NMC is impractical. Second, only a single set of Monte-Carlo trials were generated for all lag values computed, induc- ing potential correlations between the p values. These correlations should decrease as more trials are generated. Therefore results are extrapolated to NMC = 20, 000 (resulting in a minimum p ≈ 0.01) under the assumption that the processing for 20 times as many simulations would take 20 times as long. Though the bonferroni correction used here is conservative, it is less conservative (by an order of magnitude) than simulating a whole new set of spike trains for each p value as would be required to entirely eliminate any correlations between the p values. For the closed form jitter method, all lookup tables were computed de novo for each spike train. This is a conservative approach (favoring the Monte Carlo technique) since performance of the closed form jitter method could be improved by computing the tables only once and using them for all spike trains. This is certainly advised in a "production environment." The results of this simulation, shown in Figure 1, illustrate a number of features about the speed of the two algorithms. Plotted is the performance gain, defined as the ratio of the computation time between the Monte Carlo jitter method and the closed form jitter method. The first observation is that the closed form method is substantially faster than the Monte Carlo method in all cases considered. Second, while the performance gain depends only weakly on spike train length, it does decrease with increasing firing rate. This is because the computation time of the closed form jitter method increases with firing rate. In practice, however, it is rare to observe firing at sustained frequencies exceeding 100 Hz in physiological recordings. In the physiological range, the closed form jitter algorithm is faster by a factor of approximately 180 to 7200. Harrison (2013) uses importance sampling to accelerate the Monte Carlo hypothesis testing process which requires drawing fewer samples. In that work the number of samples needed, even for a low Bonferroni corrected p value, is reduced to 100. However, each sample is reported to take 18 times as long to generate and process as before, effectively resulting in a simulation about 11 times faster than the Monte Carlo simulation with NMC = 20, 000. Therefore, under physiological conditions the closed form computation has an expected speed-up of 16 to 650 times compared to the importance sampling method. It should be noted that in cases where even lower p values are needed 14 Closed Form Jitter Analysis because of multiple hypothesis constraints, importance sampling will provide larger gains in estimating very small p values. In these cases, increasing the p value requirements has no effect on the computation speed of the closed form method, so the closed form method can be expected to be faster in all cases. Another improvement mentioned in Section 2.2.2 is the ability of the closed form jitter method to compute the jitter corrected correlogram very rapidly, without computing the null hypothesis distribution of correlation values. To show the magnitude of the improvement, the simulation was repeated with only the mean of the null hypothesis distribution computed under the closed form jitter method since this is all that is needed for the corrected correlation function, eq. 5. We also restricted firing frequencies to the range 5 -- 200 Hz. Figure 2 shows the ratio of the time it takes to compute equation 2 vs. equation 5. In these cases, the closed form jitter calculation is substantially faster (480x -- 13,000x), with increasing benefits for increasing spike train lengths. As discussed previously, the spike train length is typically not that of individual trials but of the concatenation of many trials. 4 Discussion The importance, or absence of it, of precise timing of neural spikes has been discussed for the last half-century. Several techniques have been developed to characterize neuronal responses at fine time scales and it is clear that statistical methods have to be developed with much care to avoid wrong conclusions (e.g. Gawne and Richmond, 1993; Roy et al., 2000). One impor- tant difficulty is that firing rates can co-vary in the neurons under study. It is well-known that such co-variations are observable in quantities like pair- wise cross correlation functions but they are typically considered as irrelevant from the point of view of neuronal coding or of determining the connectivity in the underlying circuitry. For instance, the onset of a stimulus will typi- cally generate a temporary increase in firing rates in sensory cortex but the resulting increase in cross correlation is usually not considered of importance for neural coding (for an exception see Chase and Young, 2007, who showed that spike timing relative to onset-related population activity is informative). One common way to subtract such stimulus-locked effects is by subtracting a "shuffle predictor" (Perkel et al., 1967), obtained by computing cross cor- relations between spike trains from permuted trials. It has been pointed out 15 Closed Form Jitter Analysis repeatedly (see references in the Introduction) that this does not eliminate spurious correlations, including close to τ = 0 (synchrony). Brody (1998, 1999) and Amarasingham et al. (2012) proved that adopt- ing the null hypothesis of independent neurons can not solve the problem. Observation of such correlations is, indeed, evidence against the null hypoth- esis of independence between the two observed spike trains. Rejection of this null hypothesis can, however, occur either because spikes in the two spike trains are correlated "one-by-one" (synchrony), or because of slow firing rate covariations common to both spike trains. The fact that this null hypothesis can be rejected does not tell us why it is rejected. If the question is whether synchrony exists at less than a given time scale (only), this is the wrong null hypothesis. Instead, the time scale needs to be specified explicitly. The null hypothesis chosen by Amarasingham et al. (2012) is that changes of spike times within a time interval of size ∆ have no effect on the computed statis- tic, in this case the correlation function. It is this null hypothesis that is tested by computer simulation in the Amarasingham et al. (2012) study and analytically in this report. A key element of the methods discussed here is that the jitter intervals are defined without reference to the original spike trains. This ensures that if the null hypothesis is true, there is no way to distinguish the original spike trains from the Monte Carlo simulated spike trains. This characteristic (called exchangeability) ensures that the obtained p values are from a well formulated hypothesis test. If, on the other hand, the resampling method was changed so that each spike was jittered about it's original spike time, then even under the null hypothesis the original spike train would stand out from the rest because all of its spikes would be at the center of the jitter intervals. Therefore the resulting test would not be a proper statistical test and should be avoided (Amarasingham et al., 2012). We have discussed two ways one can choose to characterize the correla- tions between two spike trains. One is a strict hypothesis testing approach. A null hypothesis is formulated, namely that the observed correlations are indistinguishable from correlations between spike trains whose spikes have been distributed randomly within intervals of length ∆, without changing the number of spikes in each interval. By comparing the observed correla- tion with those in the distribution generated under the null hypothesis, it is then decided for a given α whether the null hypothesis can be rejected. The alternative is to compute the time-resolved correlation function and "correct" for the correlations as observed under the null hypothesis, by sub- 16 Closed Form Jitter Analysis tracting the expectation value of the latter. This is the more commonly cho- sen approach, perhaps because the time-resolved correlation function is both intuitive and familiar. The distribution of JCCG values can be compared between experimental conditions (indicating a change in 'excess synchrony') using a bootstrap test to test for significance. Also, its shape (e.g. the lo- cation of peaks) may provide insight that goes beyond the yes-no answer whether the null hypothesis can be rejected or not. In the Amarasingham et al. (2012) study, the Monte Carlo procedure is further developed to account for more potential causes of fine timing effects besides synchrony such as ramping spike rates within an interval or inter- spike interval distribution effects. These methods are straightforward and statistically well-defined. Like any Monte Carlo method, however, they only generate an approximation to the underlying distribution whose quality de- pends on the number of surrogate spike trains. In practice what is more prob- lematic is that the method can be computationally very costly. For instance, as discussed in section 3.2, our example problem using the simplest of the null hypotheses discussed (50 spike trains of a few seconds long each, mean rates between 1 and 100 Hz, maximal time lag of 100 ms, α = 0.01 with Bon- ferroni correction applied) would have required a simulation several months long on a reasonably fast machine. We therefore only simulated NM C = 1000 trials and extrapolated to the execution time needed for NM C = 20, 000 but even that abbreviated Monte Carlo run took nearly six days. Some progress can be made by using much faster machines or many machines (the problem parallelizes easily) but execution time is clearly a problem. In contrast, the closed form jitter methods this report focuses on are ex- act, rather than approximate. More important for practical applications may be that they are extremely efficient, with a speed-up of at least two orders of magnitude for the hypothesis testing approach, and four orders of magnitude for the full correlation functions. Even over importance sampling methods (Harrison, 2013), they have been shown to provide a substantial increase in speed. For the hypothesis testing examples used in our study (whose scope is quite comparable to that of typical neurophysiological experiments, assum- ing a proper Bonferroni correction is applied), computation time is reduced from more than 100 days under the original Monte Carlo method to about one night. Computational time required for the full correlation function is reduced from over 100 days to a few minutes. An increase in performance on this scale is more than merely a quantitative improvement. For instance, it is essentially impossible to explore variations in the analyses (like the influence 17 Closed Form Jitter Analysis of the jitter time scale ∆) if each computational run takes a few months, but it is easy to do if it takes minutes. So far we were only concerned with correlations between two spike trains. Modern recording techniques are already increasing the number of simulta- neously recorded spike trains to tens or hundreds. Unfortunately, the closed- form jitter method is limited in the ability to analyze large ensembles. This is because the correlation functions of some pairs in an ensemble will restrict the possible correlation values of other pairs. For example, if there are three neurons X, Y , and Z, and the pairs XY and Y Z have perfect correlation, then the pair XZ must also have perfect correlation. A Monte-Carlo jitter analysis that jitters an entire ensemble of neurons and then performs a hy- pothesis test on the ensemble can be performed relatively simply, but no such closed-form method exists yet. In order to avoid the constraints of the type described above, only N − 1 pairs of neurons can be analyzed with closed form methods when N neurons are recorded. Additionally, the nature of the exact solutions provides an opportunity for further exact analysis. Having a closed form solution allows questions about the effects of spike sorting errors, the value of ∆, or the structure of JCCG(τ ) to be addressed rigorously and more precisely than is possible with any numerical method. In conclusion, we study a statistical framework for quantifying correla- tions between spike trains at given time scales. It can be applied both for hypothesis testing and for correcting observed correlation functions for cor- relations at these time scales. Results are exact, and both computational complexity and computational time for realistic examples are several orders of magnitude lower than related approaches based on Monte Carlo simula- tions. Matlab code will be made available by the authors upon request. Acknowledgments This work was supported by the Office of Naval Research under MURI grant N000141010278 and by NIH under R01EY016281-02. We acknowledge dis- cussions with Drs. Rudiger von der Heydt and Anne Martin who also gave us access to unpublished data. 18 Closed Form Jitter Analysis References M. Abeles. Corticonics -- Neural circuits of the cerebral cortex. Cambridge University Press, 1991. E. D. Adrian and Y. Zotterman. The impulses produced by sensory nerve endings. Part 2. The response of a single end organ. J. Physiol., 61:151 -- 171, 1926. A. Amarasingham, M. T. Harrison, N. G. Hatsopoulos, and S. Geman. Con- ditional modeling and the jitter method of spike resampling. Journal of Neurophysiology, 107(2):517 -- 531, 2012. PMC3349623. C. D. Brody. Slow covariations in neuronal resting potentials can lead to arte- factually fast cross-correlations in their spike trains. J. Neurophysiology, 80(6):3345 -- 51, December 1998. C. D. Brody. Correlations without synchrony. Neural Comput, 11(7):1527 -- 1535, Oct 1999. S. M Chase and E. D Young. First-spike latency information in single neurons increases when referenced to population onset. Proceedings of the National Academy of Sciences (USA), 104(12):5175 -- 5180, 2007. J. W. Cooley and J. W. Tukey. An algorithm for the machine calculation of complex Fourier series. Mathematics of computation, 19(90):297 -- 301, 1965. T. J. Gawne and B. J. Richmond. How independent are the messages carried by adjacent inferior temporal cortical neurons? J. Neurosci., 13(7):2758 -- 2771, 1993. M. T. Harrison. Accelerated spike resampling for accurate multiple testing controls. Neural Computation, 25(2):418 -- 449, 2013. T. Hirabayashi, D. Takeuchi, K. Tamura, and Y. Miyashita. Functional mi- crocircuit recruited during retrieval of object association memory in mon- key perirhinal cortex. Neuron, 77(1):192 -- 203, January 2013a. T. Hirabayashi, D. Takeuchi, K. Tamura, and Y. Miyashita. Microcircuits for hierarchical elaboration of object coding across primate temporal areas. Science, 341(6142):191 -- 5, July 2013b. 19 Closed Form Jitter Analysis Y. Manolopoulos. Binomial coefficient computation: recursion or iteration? ACM SIGCSE Bulletin, 34(4):65 -- 67, 2002. E. Niebur, C. Koch, and C. Rosin. An oscillation-based model for the neural basis of attention. Vision Research, 33:2789 -- 2802, 1993. D. H. Perkel, G. L. Gerstein, and G. P. Moore. Neuronal spike trains and stochastic point processes. II: Simultaneous spike trains. Biophys. J., 7: 419 -- 440, 1967. A. Riehle, S. Grun, M. Diesmann, and A. Aertsen. Spike synchronization and rate modulation differentially involved in motor cortical function. Science, 278:1950 -- 1953, December 1997. A. Roy, P. N. Steinmetz, and E. Niebur. Rate limitations of unitary event analysis. Neural Computation, 12(9):2063 -- 2082, 2000. M. N. Shadlen and W. T. Newsome. The variable discharge of cortical neu- rons: implications for connectivity, computation, and information coding. J. Neurosci., 18:3870 -- 3896, 1998. W. Singer and C. M. Gray. Visual feature integration and the temporal correlation hypothesis. Annu. Rev. Neurosci., 18:555 -- 586, 1995. M. A. Smith, X. Jia, A. Zandvakili, and A. Kohn. Laminar dependence of neuronal correlations in visual cortex. Journal of Neurophysiology, 109(4): 940 -- 7, March 2013. W. Softky. Simple codes versus efficient codes. Current Opinion in Neurobi- ology, 5:239 -- 247, 1995. P. N. Steinmetz, A. Roy, P. Fitzgerald, S. S. Hsiao, K. O. Johnson, and E. Niebur. Attention modulates synchronized neuronal firing in primate somatosensory cortex. Nature, 404:187 -- 190, 2000. 20 Closed Form Jitter Analysis P value Computation 5 Hz 10 Hz 20 Hz 40 Hz 50 Hz 100 Hz 200 Hz 400 Hz 500 Hz 20 40 60 80 100 Spike Train Length [s] 4 10 3 10 i n a G e c n a m r o f r e P 2 10 1 10 0 Figure 1: Performance gain in implementing the closed form jitter method for p value computations. Gain is defined as the ratio in computation time between the Monte Carlo Jitter method and the closed form jitter method. Processing parameters used are τmax = 100ms, ∆ = 20, and NMC = 20, 000 (see text for details). 21 Closed Form Jitter Analysis JCCG Computation 4 10 5 Hz 50 Hz 100 Hz 200 Hz i n a G e c n a m r o f r e P 3 10 0 20 40 60 80 100 Spike Train Length [s] Figure 2: Performance gain in implementing the closed form jitter method for Jitter Corrected Correlogram computations. Gain is defined as the ratio in computation time between the Monte Carlo Jitter method and the closed form jitter method. Processing parameters used are τmax = 100ms, ∆ = 20, and NMC = 20, 000 (see text for details). The lowered performance gain at signal length of 1 second is due to the overhead of computing the probability table de novo for each spike train. 22
1302.3943
1
1302
2013-02-16T08:10:04
Interplay between Network Topology and Dynamics in Neural Systems
[ "q-bio.NC", "cond-mat.dis-nn", "nlin.AO", "physics.bio-ph" ]
This thesis is a compendium of research which brings together ideas from the fields of Complex Networks and Computational Neuroscience to address two questions regarding neural systems: 1) How the activity of neurons, via synaptic changes, can shape the topology of the network they form part of, and 2) How the resulting network structure, in its turn, might condition aspects of brain behaviour. Although the emphasis is on neural networks, several theoretical findings which are relevant for complex networks in general are presented -- such as a method for studying network evolution as a stochastic process, or a theory that allows for ensembles of correlated networks, and sets of dynamical elements thereon, to be treated mathematically and computationally in a model-independent manner. Some of the results are used to explain experimental data -- certain properties of brain tissue, the spontaneous emergence of correlations in all kinds of networks... -- and predictions regarding statistical aspects of the central nervous system are made. The mechanism of Cluster Reverberation is proposed to account for the near-instant storage of novel information the brain is capable of.
q-bio.NC
q-bio
Department of Electromagnetism and Physics of Matter & Institute Carlos I for Theoretical and Computational Physics, University of Granada, Spain. Interplay between Network Topology and Dynamics in Neural Systems Samuel Johnson Ph.D. Thesis Advisors: Joaqu´ın J. Torres Agudo & Joaqu´ın Marro Borau Granada, April 2011 Dedicated to my family Acknowledgements Aunque hay muchas m´as personas a las que estoy agradecido, por tantas cosas, de las que puedo enumerar en un espacio razonable, tratar´e de poner el umbral en alg´un sitio y mencionar expl´ıcitamente s´olo a aquellas que me han ayudado de alguna manera directa con esta tesis. En primer lugar quiero hacer constar mi sincero agradecimiento a mis directores: Joaqu´ın Torres no s´olo es quien me introdujo al caos, los fractales, el SOC, las redes, la simulaci´on... sino el que me mostr´o que la neurociencia es algo tan concreto y estudiable como la f´ısica estad´ıstica; Joaqu´ın Marro, por su parte, al compartir conmigo su genial perspectiva sobre la ciencia y la complejidad, me ha abierto los o jos a toda una manera de ver el mundo. Miguel ´Angel Munoz ha sido tambi´en una influencia cientf´ıfica important´ısima, con su maravillosa combinaci´on de talento y buen humor. Agradezco tambi´en a tod@s l@s dem´as companer@s con quienes he traba jado y convivido estos anos, tanto por lo que he aprendido de ellos como por los muchos buenos ratos: Pablo Hurtado, Pedro Garrido, Paco de los Santos, Antonio Lacomba, Elvira Romera, Ram´on Contreras, Paco Ramos, Jes´us Cort´es, Juan Antonio Bonachela, Luca Donetti, Jos´e Manuel Mart´ın, Omar al-Hammal, Jes´us del Pozo, Carlos Espigares, Clara Guglieri, Pablo Sartori, Marina Manrique, Jordi Hidalgo, Virginia Dom´ınguez, Jorge Mej´ıas, Sebastiano de Franciscis, Alejandro Pinto, Leticia Rubio, Jordi Garces, Luca Sabino, Simone Pigolotti, Lu´ıs Seoane, Daniele Vilone y Miguel Ib´anez. Como bien sab´eis, sois companer@s mucho m´as que de traba jo, si´endolo tambi´en, seg´un el caso, de piso, de grupo musical, de decrepitud, de juegos diversos, del alma... Por supuesto, hay otras personas con esta multiplicidad de roles a quienes sin embargo no voy a tratar de nombrar aqu´ı: por favor, no pens´eis que es por falta de reconocimiento, sino por ahorrar papel y tinta, y porque de todas maneras seguramente no vay´ais a leer esto. I’m grateful to Mike Ramsey, Mohammed Boudjada and Helmut Rucker for treating me so well in Graz as well as introducing me to the world of research. Doy gracias a Marcelo del Castillo por invitarme a la UNAM y por ser, junto v vi con su familia y amig@s, tan buen anfitri´on en M´exico; a Ezequiel Albano, Gabriel Baglietto, Bel´en Moglia, Nara Guisoni y Luis Diambra por tratarme tan bien en La Plata; and to Nick Jones and Sumeet Agarwal for making me so welcome in Oxford. I am also grateful to all the people who have helped in one way or another with my research, be it reading manuscripts, providing data, suggesting ideas, or simply with stimulating conversations. I’m proba- bly leaving many deserving people out when I mention Dante Chialvo, ´Alex Arenas, Ginestra Bianconi, Yamir Moreno, Jennifer Dunne, Alberto Pascual, V´ıctor Egu´ıluz, Sasha Goltsev, Gorka Zamora, Lars Rudolf, Sabine Hilfiker, Tiago Peixoto, Ole Paulsen and Peter Latham. Tambi´en estoy agradecido a Juan Soler, Alberto Prieto, Juan Calvo, Pilar Guerrero, Irene Mendoza, Es- trella Ryan, Luna ´Alvarez, Javier (Chancly) Pascual, Nikolina Dimitrov, Felisa Torralba y Caroline de Cannart por su diversa ayuda. Three members of my family who have been particularly influential on my scientific interests and helpful in various ways are my uncle Dave Jones, my grandfather Tony Jones, and my aunt Sue Ziebland. C´ecile Poirier, qui as ´et´e avec moi pendant la plus part de ces derni`eres quatre ans, tu m’as fait surmonter tant d’obstacles et m’as aid´e tellement que je n’ai vraiment pas de mots pour te remercier assez. And the mental freedom and emotional basis I need to undertake this or any other pro ject is grounded in the unconditional support and love from my parents, Jenni and Alan, and my sisters, Jazz and Abby. Thanks so much to all of you. Finally, thanks are also due to the government of the United States for its perseverance in attempting to crush the news organization Wikileaks, whose continual publication of enlightening documents so often kept me from work- ing on this thesis; and to myself, for at all times resisting the temptation of perfectionism when writing this up – as the reader will no doubt observe. “To say that a man is made up of certain chemical elements is a satisfactory description only for those who intend to use him as a fertilizer.” Hermann Joseph Muller “Je n’avais pas besoin de cette hypoth`ese-l`a.” Pierre-Simon Laplace “Research is what I’m doing when I don’t know what I’m doing.” Werner von Braun Abstract This thesis is a compendium of research which brings together ideas from the fields of Complex Networks and Computational Neuroscience to address two questions regarding neural systems: 1) How the activity of neurons, via synaptic changes, can shape the topology of the network they form part of, and 2) How the resulting network structure, in its turn, might condition aspects of brain behaviour. Although the emphasis is on neural networks, several theoretical findings which are relevant for complex networks in general are presented – such as a method for studying network evolution as a stochastic process, or a theory that allows for ensembles of correlated networks, and sets of dynamical elements thereon, to be treated mathematically and computationally in a model-independent manner. Some of the results are used to explain experimental data – cer- tain properties of brain tissue, the spontaneous emergence of correlations in all kinds of networks... – and predictions regarding statistical aspects of the central nervous system are made. The mechanism of Cluster Reverberation is proposed to account for the near-instant storage of novel information the brain is capable of. ix Preamble: The Ant, the Grasshopper and Complexity Once upon a time, in a charming and peaceful little valley, a grasshopper sat under the shade of a sunflower, idly strumming up a tune, when a young worker ant came into view. The grasshopper watched as she trundled her way laboriously up an incline under the weight of a large piece of leaf. When she was close enough, he hailed her: ‘Ahoy there, friend. I hope I won’t seem tactless if I point out what a sin- gularly cumbersome bit of leaf you have there. Would you not rather put it down for a while and join me for a quick jam session? You could bang along on some twigs or something.’ ‘Thank you for the offer, but I must continue on my way,’ replied the ant, glancing up in slight surprise at being thus addressed. ‘Oh, what a pity,’ the grasshopper rejoined. ‘And where, if I may be so bold as to inquire, would you be taking your rather unappetising ration of cellulose?’ ‘Well, I can’t say I really know... I just follow this trail of pheromones I’ve come across. I’m sure it’s for some noble purpose though.’ ‘Ah, that must be reassuring. And I suppose when you get to wherever it is you don’t know you’re going you intend to eat your bit of leaf...’ ‘Oh no, I can’t digest something like this – who do you take me for?’ ‘You can’t? Well, how strange...’ ‘What’s strange?’ ‘However did an animal evolve which, instead of engaging in biologically rea- sonable (not to mention enjoyable) activities, such as playing music to attract sexual partners, prefers to lug useless bits of leaf about? How on earth can that serve to spread your genes?’ ‘I’m not interested in music or sex, whatever those are. I just follow simple rules, like all my identical sisters. You could say we’re automata.’ ‘Thanks, I was going to but wasn’t sure whether you’d be offended. Well, let xi xii me wish you an agreeable day of toil, you frigid little automaton.’ With that, the grasshopper gave a big leap into the air, slightly exasper- ated by the folly so often displayed by his fellow insects. Looking down, he spotted a few more ants, all carrying leaves in the same direction as the one he had just met. Intrigued, he fluttered slightly higher (since grasshoppers can, actually, fly, if not all that well). He realised the ants were all heading for a nest some way off. In fact, there were many ant trails leading to various sources of food. It dawned on the grasshopper that although the individual ants were just boring little morons idiotically following rules, the nest as a whole was managing to find the closest leaves, bring them back along optimal routes, and feed them to its plantations of fungi. The colony was behaving like an intelligent organism, in some respects not so different from he himself, who functioned thanks to the cells of his body – each with the same genome, like the ants – cooperating through the obedience to relatively simple rules. This thought impressed the grasshopper very much, driving him to flutter even higher so as to see things in greater perspective. From there he considered the apparently fragile web of trophic, parasitical and symbiotic interactions linking all the living beings in the valley – a network which nonetheless must have evolved a particularly robust structure not to shatter at the first environ- mental fluctuation. He became so enthralled by the idea of such complexity on one scale emerging from simplicity on another that he didn’t even pay any attention to an attractive young grasshopperess making her wanton way just below him. Instead, he couldn’t help fearing that a butterfly he noticed gently flapping his wings would probably set off a hurricane somewhere. As he flew ever higher, he began to see snowflakes glide by, overwhelmingly intricate and beautiful patterns self-organised out of the simplest little water molecules. Fi- nally he was so high that he began to reflect on how the very stellar systems, galaxies, clusters, superclusters, filaments of galaxies... – of which his whole world was but an infinitesimal component – also interacted with each other via the simple rules of gravity and pressure to form ob jects marvellous beyond conception. What he didn’t notice until it was too late, as he left behind the cosy protection of the atmosphere, was how ultraviolet sunlight and ionising cos- mic rays were steadily burning his wings each to a crisp. Beginning to fall, he xiii only hoped he would have time to consider the several morals to his tragic tale. After a while spent plummeting to his doom he realised that, the freefall terminal velocity and life expectancy of a grasshopper being what they respec- tively were, he would most likely die peacefully of old age somewhere along his way down – never again contemplating his Edenic valley except, like some prophetic locust, from afar. Contents Acknowledgements Abstract Preamble: The Ant, the Grasshopper and Complexity List of figures List of tables 1 Resumen en espanol v ix xi xxv xxvii 1 7 2 Where we are and where we’d like to go 2.1 From bridges to brains . . . . . . . . . . . . . . . . . . . . . . . 7 2.2 Neural networks in neuroscience . . . . . . . . . . . . . . . . . . 13 2.3 A declaration of intent . . . . . . . . . . . . . . . . . . . . . . . 18 3 Evolving networks and the development of neural systems 21 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 3.1 3.2 Basic considerations . . . . . . . . . . . . . . . . . . . . . . . . 23 3.3 Synaptic pruning . . . . . . . . . . . . . . . . . . . . . . . . . . 25 3.4 Phase transitions . . . . . . . . . . . . . . . . . . . . . . . . . . 26 3.5 Correlations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3.6 The C. Elegans neural network . . . . . . . . . . . . . . . . . . 34 3.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 39 4 Bringing on the Edge of Chaos with heterogeneity 4.1 Exciting cooperation . . . . . . . . . . . . . . . . . . . . . . . . 39 4.2 The Fast-Noise model . . . . . . . . . . . . . . . . . . . . . . . . 41 4.3 Edge of Chaos . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 4.4 Network performance . . . . . . . . . . . . . . . . . . . . . . . . 47 xiv Contents xv 4.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 5 Correlated networks and natural disassortativity 51 5.1 Assortativity of networks . . . . . . . . . . . . . . . . . . . . . . 51 5.2 The entropy of network ensembles . . . . . . . . . . . . . . . . . 53 5.3 Entropic origin of disassortativity . . . . . . . . . . . . . . . . . 55 5.4 To sum up... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 6 Enhancing robustness to noise via assortativity 61 6.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 6.2 Preliminary considerations . . . . . . . . . . . . . . . . . . . . . 64 6.2.1 Model neurons on networks . . . . . . . . . . . . . . . . 64 6.2.2 Network ensembles . . . . . . . . . . . . . . . . . . . . . 65 6.2.3 Correlated networks . . . . . . . . . . . . . . . . . . . . 66 6.3 Analysis and results . . . . . . . . . . . . . . . . . . . . . . . . . 68 6.3.1 Mean field . . . . . . . . . . . . . . . . . . . . . . . . . . 68 6.3.2 Generating correlated networks . . . . . . . . . . . . . . 71 6.3.3 Assortativity and dynamics . . . . . . . . . . . . . . . . 72 6.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 7 Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning 79 7.1 Slow but sure, or fast and fleeting? . . . . . . . . . . . . . . . . 79 7.2 The simplest neurons on modular networks . . . . . . . . . . . . 83 7.3 Cluster Reverberation . . . . . . . . . . . . . . . . . . . . . . . 84 7.4 Energy and topology . . . . . . . . . . . . . . . . . . . . . . . . 88 7.5 Forgetting avalanches . . . . . . . . . . . . . . . . . . . . . . . . 88 7.6 Clustered networks . . . . . . . . . . . . . . . . . . . . . . . . . 90 7.7 Yes, but does it happen in the brain? . . . . . . . . . . . . . . . 93 8 Concluding remarks 9 Conclusiones en espanol 94 98 A Nonlinear preferential rewiring in fixed-size networks as a dif- fusion process 102 B Effective modularity of highly clustered networks 110 xvi Contents C Nestedness of networks 112 C.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 C.2 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 C.3 The effect of the degree distribution . . . . . . . . . . . . . . . . 115 C.4 Nestedness and assortativity . . . . . . . . . . . . . . . . . . . . 116 C.5 Bipartite networks . . . . . . . . . . . . . . . . . . . . . . . . . 118 C.6 Overlapping networks . . . . . . . . . . . . . . . . . . . . . . . . 120 C.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 D Publications derived from the thesis D.1 Journals and book chapters (the most relevant ones marked with . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 an asterisk) D.2 Abstracts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 124 References 129 List of Figures 2.1 The problem of the Seven Bridges of Konigsberg can be reduced to a graph in which nodes and edges represent land masses and . . . . . . . . . . . . . . . . . . . . . . . . bridges, respectively. 8 2.2 Drawing of the cells of the chick cerebellum, from “Estructura de los centros nerviosos de las aves”, Madrid, 1905. Notice how the neurons make up a complex network of synaptic interactions. 14 2.3 In the Hopfield neural-network model, the interaction strengths (representing synaptic weights) store information in the form of particular patterns, or configurations of firing neurons, which become attractors of the dynamics. Whatever the initial state of the system, it will always evolve towards one of these patterns, thus allowing for the storage and retrieval of information. The mechanism, known as associative memory, is thought to be at the basis of memory in the brain. In this case, the network is “remembering” an illustration by Jean-Baptiste Oudry for Jean . . . . . . . . . . 16 de la Fontaine’s fable La Cigale et la Fourmi. xviii List of Figures xix 3.1 Synaptic densities in layers 1 (red squares) and 2 (black circles) of the human auditory cortex against time from conception. Data from Huttenlocher and Dabholkar (1997), obtained by di- rectly counting synapses in tissues from autopsies. Lines follow best fits to Eq. (3.7), where the parameters were: for layer 1, τp = 5041 days; and for layer 2, τp = 3898 days (for ρf we have used the last data pints: 30.7 and 40.8 synapses/µm3 , for layers 1 and 2 respectively). Data pertaining to the first year and to days 4700, 5000 7300, shown with smaller symbols, where omit- ted from the fit. Assuming the existence of transient growth factors, we can include the data points for the first year in the fit by using Eq. (3.8). This is done in the inset (where time is displayed logarithmically). The best fits were: for layer 1, τg = 151.0 and τp = 5221; and for layer 2, τg = 191.1 and τp = 4184, all in days (we have approximated t0 to the time of the first data points, 192 days). . . . . . . . . . . . . . . . . . . 27 3.2 Evolution of the degree distributions of networks beginning as regular random graphs with κ(0) = 20 in the critical (top) and supercritical (bottom) regimes. Local probabilities are σ(k) = k/(hkiN ) in both cases, and π(k) = 2σ(k) − N −1 and π(k) = k3/2 /(hk3/2iN ) for the critical and supercritical ones, respec- tively. Global probabilities as in Eq. (3.6), with n = 10 and κmax = 20. Symbols in the main panels correspond to p(k , t) at different times as obtained from MC simulations. Lines result from numerical integration of Eq. Insets show typical (3.1). time series of κ and m. Light blue lines are from MC simula- tions and red lines are theoretical, given by Eq. (3.7) and Eq. . . . . . . . . . . . . . . . . . . . 28 (3.1), respectively. N = 1000. xx List of Figures 3.3 Phase transitions in mst for π(k) ∼ kα and σ(k) ∼ k , and u(κ) and d(κ) as in Eq. (3.6). N = 1000 (blue squares), 1500 (red triangles) and 2000 (black circles); κ(0) = κmax = 2n = N/50. Corresponding lines are from numerical integration of Eq. (3.1). The bottom left inset shows values of the highest eigenvalue of the Laplacian matrix (red squares) and of Q = λN /λ2 (black circles), a measure of unsynchronizability; N = 1000. The top right inset shows transitions for the same parameters in the final values of Pearson’s correlation coefficient r (see Section 3.5), both for only one edge allowed per pair of nodes (red squares) and without this restriction (black circles). . . . . . . . . . . . 29 3.4 Degree distribution (binned) of the C. Elegans neural network (circles) (White et al., 1986) and that obtained with MC simu- lations (line) in the stationary state (t = 105 steps) for an equiv- alent network in which edges are removed randomly (β = 1) at the critical point (α = 1.35). N = 307, κst = 14.0, averages over 100 runs. Global probabilities as in Eq. (3.6). The slope is for k−5/2 . Top right inset: mean-neighbour-degree function knn (k) as measured in the same empirical network (circles) and as given by the same simulations (line) as in the main panel. The slope is for k−1/2 . Bottom left inset: mst of equivalent net- work for a range of α, both from simulations (circles) and as obtained with Eq. (3.1). (See also Table 3.1.) . . . . . . . . . . 35 4.1 The temperature dependence of the difference between the val- ues for the fatigue at which the ferromagnetic–periodic transi- tion occurs, as obtained analytically for T = 0 (Φ0 ) and from MC simulations at finite T (Φc ). The critical temperature is cal- culated as Tc = hk2i (hki N )−1 for each topology. Data are for bimodal distributions with varying ∆ and for scale–free topolo- gies with varying γ , as indicated. Here, hki = 20, N = 1600 and α = 2. Standard deviations, represented as bars in this graph, were shown to drop with N −1/2 (not depicted). . . . . . . . . . 45 List of Figures xxi 4.2 The critical fatigue values Φ0 (solid lines) and Φc from MC av- erages over 10 networks (symbols) with T = 2/N , hki = 20, N = 1600, α = 2. The dots below the lines correspond to changes of sign of the Lyapunov exponent as given by the iter- ated map, which qualitatively agree with the other results. This . . . . . . 46 is for bimodal and scale–free topologies, as indicated. 4.3 Network “performance” (see the main text) against ∆ for bi- modal topologies (above) and against γ for scale–free topologies (below). Φ = 0.8 for the first case and Φ = 1 in the second. Averages over 20 network realizations with stimulation every 50 MC steps for 2000 MC steps, δ = 5 and M = 4; other param- eters as in Fig. 6.5. Inset shows sections of typical time series of mν for ∆ = 10 (above) and γ = 4 (below); the corresponding stimulus for pattern ν is shown underneath. . . . . . . . . . . . 48 5.1 Evolution of the Internet at the AS level. Empty (blue) squares and circles: entropy per node of randomized networks in the fully random and in the configuration ensembles, as obtained by Bianconi (hence the “B” superscription) (Bianconi, 2008, 2009; Anand and Bianconi, 2009). Filled (red) triangles and diamonds: Shannon entropy for an ER network and a scale-free one with γ = 2.3, respectively. . . . . . . . . . . . . . . . . . . 55 5.2 Shannon entropy of correlated scale-free networks against pa- rameter β (left panel) and against Pearson’s coefficient r (right panel), for various values of γ (increasing from bottom to top). hki = 10, N = 104 . . . . . . . . . . . . . . . . . . . . . . . . . . 56 5.3 Lines from top to bottom: r at which the entropy is maxi- mized, r∗ , against γ for random scale-free networks with mean degrees hki = 1 2 , 1, 2 and 4 times k0 = 5.981, and N = N0 = 10697 nodes (k0 and N0 correspond to the values for the Inter- net at the AS level in 2001 (Park and Newman, 2003), which had r = r0 = −0.189). Symbols are the values obtained in (Park and Newman, 2003) as those expected solely due to the one-edge-per-pair restriction (with k0 , N0 and γ = 2.1, 2.3 and 2.5). Inset: r∗ against N for networks with fixed hki/N (same values as the main panel) and γ = 2.5; the arrow indicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 N = N0 . xxii List of Figures 5.4 Level of assortativity that maximizes the entropy, r∗ , for vari- ous real-world, scale-free networks, as predicted theoretically by Eq. (5.1) (circles) and as directly measured (horizontal lines), against exponent γ . . . . . . . . . . . . . . . . . . . . . . . . . 60 6.1 Mean-nearest-neighbour functions knn (k) for scale-free networks with β = −0.5 (disassortative), 0.0 (neutral), and 0.5 assor- tative, generated according to the algorithm described in Sec. 6.3.2. Inset: degree distribution (the same in all three cases). Other parameters are γ = 2.5, hki = 12.5, N = 104 . . . . . . . . 66 6.2 Stable stationary value of the weighted overlap µ1 against tem- perature T for scale-free networks with correlations according to knn ∼ kβ , for β = −0.5 (disassortative), 0.0 (neutral), and 0.5 (assortative). Symbols from MC simulations, with errorbars representing standard deviations, and lines from Eqs. (6.6). Other network parameters as in Fig. 6.1. Inset: µ1 against T for the assortative case (β = 0.5) and different system sizes: N = 104 , 3 · 104 and 5 · 104 . . . . . . . . . . . . . . . . . . . . . 72 6.3 Stable stationary values of order parameters µ0 , µ1 and µβ+1 against temperature T , for assortative networks according to β = 0.5. Symbols from MC simulations, with errorbars repre- senting standard deviations, and lines from Eqs. (6.6). Other parameters as in Fig. 6.1. . . . . . . . . . . . . . . . . . . . . . 73 6.4 Difference between the stationary values µ1 and µ0 for networks with β = −0.5 (disassortative), 0.0 (neutral) and 0.5 (assorta- tive), against temperature. Symbols from MC simulations, with errorbars representing standard deviations, and lines from Eqs. (6.6). Line shows the expected level of fluctuations due to noise, ∼ N − 1 2 . Other parameters as in Fig. 6.1. . . . . . . . . . . . . 73 6.5 Phase diagrams for scale-free networks with γ = 2.5, 3, and 3.5. Lines show the critical temperature Tc marking the second-order transition from a memory (ferromagnetic) phase to a memory- less (paramagnetic) one, against the assortativity β , as given by Eq. (6.7). Other parameters as in Fig. 6.1. . . . . . . . . . . . 75 6.6 Parameter space β − γ partitioned into the regions in which b(β , γ ) has the same functional form – where b is the scaling exponent of the critical temperature: Tc ∼ N b . Exponents ob- . . . . . . . . . 75 tained by taking the large N limit in Eq. (6.7). List of Figures xxiii 6.7 Examples of how Tc scales with N for networks belonging to regions I, II, III and IV of Fig. 6.6 (β = −0.8, −0.35, 0.0 and 0.9, respectively). Symbols from MC simulations, with errorbars representing standard deviations, and slopes from Eq. (6.7). All parameters – except for β and N – are as in Fig. 6.1. . . . . . 76 6.8 Global order parameter ζ for assortative (β = 0.5), neutral (β = 0.0) and disassortative (β = −0.5) networks with P = 3 (left panel) and P = 10 (right panel) stored patterns. Symbols from MC simulations, with errorbars representing standard de- viations. All parameters are as in Fig. 6.1. . . . . . . . . . . . 77 7.1 Diagram of a modular network composed of four five-neuron clusters. The four circles enclosed by the dashed line represent the stimulus: each is connected to a particular module, which adopts the input state (red or blue) and retains it after the stimulus has disappeared via Cluster Reverberation. . . . . . . . 85 7.2 Performance η against λ for networks of the sort described in the main text with M = 160 modules of n = 10 neurons, hki = 9; patterns are shown with intensities δ = 8.5, 9 and 10, and T = 0.02 (lines – splines – are drawn as a guide to the eye). Inset: typical time series of mstim (i.e., the overlap with whichever pattern was last shown) for λ = 0.0, 0.25, and 0.5, and δ = hki = 9. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 7.3 Configurational energy of a network composed of M = 20 mod- ules of n = 10 neurons each, according to Eq. (7.1), for various values of the rewiring probability λ. The minima correspond to situations such that all neurons within any given module have the same sign. . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 7.4 Left panel: distribution of escape times τ , as defined in the main text, for λ = 0.25 and T = 2. Slope is for β = 1.35. Other parameters as in Fig. 7.2. Symbols from MC simulations and line given by Eqs. (7.2) and (7.3). Right panel: exponent β of the quasi-power-law distribution p(τ ) as given by Eq. (7.3) for temperatures T = 1 (red line), T = 2 (green line) and T = 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 (blue line). xxiv List of Figures 7.5 Proportion of outgoing edges, λ, from boxes of linear size l against exponent γ for scale-free networks embedded on 2D lat- tices. Lines from Eq. (7.4) and symbols from simulations with hki = 4 and N = 1600. . . . . . . . . . . . . . . . . . . . . . . 92 7.6 Performance η against exponent γ for scale-free networks, em- bedded on a 2D lattice, with patterns of M = 16 modules of n = 100 neurons each, hki = 4 and N = 1600; patterns are shown with intensities δ = 3.5, 4, 5 and 10, and T = 0.01 (lines – splines – are drawn as a guide to the eye). Inset: typical time series for γ = 2, 3, and 4, with δ = 5. . . . . . . . . . . . . . . . 92 A.1 Degree distribution p(k , t) at four different stages of evolution: t = 102 [(yellow) squares], 103 [(blue) circles], 104 [(red) trian- gles)] and 105 MCS [(black) diamonds]. From top to bottom panels, subcritical (α = 0.5), critical (α = 1) and supercritical (α = 1.5) rewiring exponents. Symbols from MC simulations and corresponding solid lines from numerical integration of Eq. . . . . . . . . 106 (A.1). β = 1, hki = 10 and N = 1000 in all cases. A.2 Adjusted variance σ 2/hki2 of the degree distribution after 2×105 MCS against α, as obtained from MC simulations, for system sizes N = 800 [(yellow) squares], 1200 [(blue) circles], 1600 [(red) triangles] and 2000 [(black) diamonds]. Top left inset shows final degree distributions for α = 0.5 [light gray (blue)], 1 [dark gray (red)] and 1.5 (black), with N = 1000. Bottom right inset shows typical time series of σ 2/hki2 for the same three values of α and N = 1200. In all cases, β = 1 and hki = 10. . . . 107 C.1 Maximally packed matrix representing a network of plants and islands off Perth (Abbott and Black, 1980) (because the net- work is bipartite, the adjacency matrix is composed of four blocks: two identical to this matrix, the other two composed of zeros). Data, image and line obtained from NESTEDNESS CALCULATOR, which returns a “temperature” of T = 0.69o for this particular network. . . . . . . . . . . . . . . . . . . . . 113 C.2 Nestedness against assortativity (as measured by Pearson’s cor- relation coefficient) for scale-free networks as given by Eq. (C.13). . . . . . . . . . . . . . . . . . . . . . . . . 119 hki = 10, N = 1000. List of Figures xxv C.3 Nestedness against assortativity (as measured by Pearson’s cor- relation coefficient) for data on a variety of networks. Blue squares are food webs (Table C.4) and red circles are networks of all other types (Table C.4). . . . . . . . . . . . . . . . . . . . 119 List of Tables 3.1 Values of small-world parameters C and l, and Pearson’s cor- relation coefficient r , as measured in the neural network of the worm C. Elegans (White et al., 1986), and as obtained from sim- ulations in the stationary state (t = 105 steps) for an equivalent network at the critical point when edges are removed randomly – i.e., for α = 1.35 and β = 1. N = 307, κst = 14.0; averages over 100 runs and global probabilities as in Eq. (3.6). The- oretical estimates correspond to Eqs. (3.12), (3.14) and (3.11) applied to the networks generated by the same simulations. The last column lists the respective configuration model values: C and l are obtained theoretically as in (Newman, 2003c), while r , from MC simulations as in (Maslov et al., 2004), is the value expected due to the absence of multiple edges. (See also Fig. 3.4.) 36 C.1 Food webs appearing in Fig. C.3 (listed from least to most assortative) : r is the assortativity and ν the nestedness. The origins of all data cited in Ref. (Dunne et al., 2004), and kindly provided to us by Jennifer Dunne. . . . . . . . . . . . . . . . . 120 C.2 Empirical networks appearing in Fig. C.3 (listed from least to most assortative) : r is the assortativity and ν the nestedness. All data available on the personal Web pages of ´Alex Arenas, . . . . . . . . . . . . . . . . 121 Mark Newman and Duncan Watts. xxvii Chapter 1 Resumen en espanol Paradigma de sistema complejo y el peor comprendido de nuestros ´organos, el cerebro es, esencialmente, una inmensa red de c´elulas que se comunican entre s´ı mediante senales electro-qu´ımicas. Este traba jo recoge y desarrolla ideas del joven campo de las Redes Complejas para tratar de mejorar nuestro en- tendimiento acerca del comportamiento colectivo complejo que puede emerger en las redes de neuronas a partir de din´amicas individuales relativamente sen- cillas. El Cap´ıtulo 2 es una breve introducci´on a las Redes Complejas y a la Neurociencia Computacional. Se describe, entre otras cosas, el modelo de Hopfield de red neuronal atractora, en que cada nodo representa una neurona y las sinapsis son representadas por los enlaces. Este sistema puede almacenar informaci´on, en forma de patrones o configuraciones concretas de neuronas activas e inactivas, en los pesos sin´apticos; es decir, en la intensidad con la que la actividad de una neurona influye sobre sus vecinas. Si, para representar un patr´on dado, dos neuronas vecinas han de adoptar el mismo estado (activo o inactivo), se refuerza la interacci´on entre ambas, mientras que se debilita en caso contrario. Repitiendo esta operaci´on para cada pareja conectada de neuronas y para cada patr´on, estos patrones se convierten en los estados que minimizan la energ´ıa total (atractores de la din´amica), y el sistema evolu- ciona siempre hacia el patr´on que m´as se parezca a su estado inicial. Este mecanismo, llamado de memoria asociativa, es la responsable del almacena je y la recuperaci´on de informaci´on tanto en modelos m´as realistas de medios neuronales, como en muchos aparatos artificiales que desempenan tareas tales como la identificaci´on y la clasificaci´on de im´agenes. Adem´as, hoy en d´ıa existen evidencias experimentales suficientes para asegurar que algo parecido ocurre en el cerebro: mediante los procesos bioqu´ımicos de potenciaci´on de 1 2 Chapter 1. Resumen en espanol largo plazo (LTP, por sus siglas en ingl´es) y depresi´on de largo plazo (LTD), las sinapsis modifican gradualmente sus conductancias durante el aprendiza je. El Cap´ıtulo 3 aborda el problema de c´omo puede desarrollarse una red con el tipo de estructuras que se observa en el cerebro. Para ello se formaliza como un proceso estoc´astico una red que evoluciona mediante cambios probabil´ısticos que dependen de cualquier manera de informaci´on local y global de los grados (n´umeros de vecinos) de los nodos, tal como se hace en la Ref. (Johnson et al., 2010a). Se considera que estas suposiciones son relevantes para el caso del cerebro ya que la arborizaci´on y la atrofia sin´apticas dependen de la actividad el´ectrica de la neurona en cuesti´on, que a su vez puede estar relacionada con el n´umero de vecinos que tenga, y con la densidad sin´aptica media en la red. Se demuestra c´omo esta situaci´on viene descrita por una ecuaci´on de Fokker- Planck, y se aplica a dos conjuntos de datos reales neurofisiol´ogicos: por una parte, la curvas de poda sin´aptica (fuerte reducci´on de la densidad sin´aptica que sufre el c´ortex durante la infancia) de autopsias humanas pueden explicarse con unas suposiciones m´ınimas; por otra, varias magnitudes estad´ısticas de la red del an´elido C. Elegans (distribuci´on de grados, perfil de correlaciones, clus- tering o agrupamiento y camino m´ınimo medio) emergen con cierta precisi´on y de manera natural justo en la transici´on de fase que presenta el modelo. Esto da fuerza a la idea de que el sistema nervioso optimiza su rendimiento coloc´andose cerca de un punto cr´ıtico. Un caso parecido, en que los enlaces de la red, en vez de desaparecer o aparecer, son redirigidos estoc´asticamente, presentado en la Ref. (Johnson et al., 2009b), se describe en el Ap´endice A. El resto de la tesis se centra en los efectos que pueden tener sobre el com- portamiento colectivo de sistemas de neuronas las caracter´ısticas topol´ogicas descritas en el Cap´ıtulo 3. Se sabe que la heterogeneidad de la distribuci´on de grados de la red suele tener una influencia significativa en la din´amica de elementos conectados mediante sus enlaces. En el caso de redes neuronales de Hopfield, Torres et al. (Torres et al., 2004) demostraron que, en redes libres de escala (que son altamente heterog´eneas), el rendimiento aumenta con la heterogeneidad. El Cap´ıtulo 4 examina el mismo efecto en una red neuronal que incluye otro ingrediente biol´ogico: la depresi´on sin´aptica, gracias a la cual se observa una transici´on entre una fase de memoria est´atica a otra en que el sistema salta ca´oticamente entre los patrones guardados. Resulta que cerca de este punto cr´ıtico (el famoso Borde del Caos) la red es capaz de realizar una tarea din´amica necesaria para los seres vivos: reconocer, de entre varios patrones que tenga almacenados, uno dado que se le “ensene” y retenerlo in- 3 definidamente despu´es. Como demostramos en la Ref. (Johnson et al., 2008), la heterogeneidad de la distribuci´on de grados de la red acerca el punto cr´ıtico a una regi´on del espacio de par´ametros en que apenas hay depresi´on sin´aptica. Teniendo en cuenta que esta depresi´on empeora la capacidad de memoria del sistema, se concluye que una red altamente heterog´enea es la ´optima para re- alizar este tipo de tarea din´amica. Las redes funcionales medidas en el c´ortex humano durante tareas del estilo adopta la red libre de escala m´as heterog´ena posible, por lo que cabe la hip´otesis de que el cerebro est´e maximizando as´ı su rendimiento. Otra propiedad altamente estudiada de las redes complejas es la existencia de correlaciones entre los grados de nodos vecinos. Cuando dichas correla- ciones son positivas (nodos muy conectados se suelen conectar con otros que tambi´en tienen muchos vecinos, y los que tienen pocos con otros parecidos) se dice que la red es asortativa; mientras que es disasortativa si las correlaciones son negativas (los que tienen muchas conexiones se conectan, preponderante- mente, con los que tienen pocas). Curiosamente, se hab´ıa observado que por lo general las redes sociales (por ejemplo, redes de colaboraciones profesionales o de contactos sexuales) son asortativas, mientras que pr´acticamente todas las dem´as (gen´eticas, tr´oficas, proteicas, de transportes, de palabras, Internet, la Web...) son significativamente disasortativas. Aunque se hab´ıa estudiado los efectos de estas correlaciones en varios sistemas, las t´ecnicas matem´aticas y computacionales para ello padec´ıan de inconvenientes que restaban general- idad a los resultados. Para solventar esto, en el Cap´ıtulo 5 se describe un nuevo m´etodo para particionar el espacio de las fases de redes en regiones de correlaciones iguales, una t´ecnica que permite tanto an´alisis te´orico como com- putacional de este tipo de sistemas. Utilizando este m´etodo junto con ideas de Teor´ıa de la Informaci´on se demuestra tambi´en el resultado principal de la Ref. (Johnson et al., 2010b): que la disasortatividad es el estado “natural” (en cuanto a situaci´on de equlibrio) de las redes heterog´eneas, lo cual explica la preponderancia en la realidad de este tipo de configuraciones. La prefer- encia de los humanos por agregarse en funci´on de propiedades similares ser´ıa la explicaci´on de que las redes sociales se encuentren fuera del equilibrio, en regiones asortativas del espacio de fases. En el Cap´ıtulo 6 se aplica el m´etodo del Cap´ıtulo 5 al caso de una red neu- ronal de Hopfield que no s´olo presenta heterogeneidad, sino tambi´en correla- ciones nodo-nodo. Se encuentra, como ya fue descrito en la Ref. (de Franciscis et al., 2011), que estos sistemas pueden aumentar de manera notable su robustez 4 Chapter 1. Resumen en espanol frente a ruido gracias a las correlaciones positivas. De nuevo, esto parece enca jar, al menos cualitativamente, con resultados experimentales que han encontrado redes funcionales en el c´ortex humano altamente asortativas. Hemos dicho que las redes neuronales pueden aprender gracias a una apropi- ada modificaci´on de los pesos sin´apticos mediante LTP y LTD, lo que explica la memoria de largo plazo. Pero estos procesos bioqu´ımicos ocurren en un tiempo caracter´ıstico de al menos minutos. Los modelos de memoria de corto plazo, que ocurren en escalas de tiempo menores, suelen dar por hecho que la informaci´on que se utiliza est´a de antemano almacenada en el cerebro, y que el sujeto realizando la tarea s´olo ha de activar y mantener de alguna manera la configuraci´on correcta (como en el Cap´ıtulo 4). Pero es f´acil darse cuenta de que esto no puede ser el caso para cualquier tarea: basta mirar algo totalmente nuevo por un instante, cerrar los o jos, y pensar en lo que se ha visto. Los ´unicos modelos de memoria de corto plazo existentes que no requieren aprendiza je sin´aptico se basan en que cada neurona mantenga de alguna manera la informaci´on que le corresponde (t´ıpicamente gracias a una serie de procesos sub-celulares). Pero al no emerger la memoria como propiedad colectiva del sistema, sino como suma de memorias individuales, estos modelos padecen de una gran falta de robustez frente a ruido. Y, lejos de presentar un comportamiento individual fiable, las neuronas se caracterizan justamente por ser c´elulas de una alta variabilidad, con tendencia a disparar de manera m´as o menos aleatoria. En el Cap´ıtulo 7 se propone un mecanismo, llamado Cluster Reverberation (CR), o Reverberaci´on de Grupo, gracias al cual incluso sistemas como redes con unidades simples, binarias, como en el modelo de Hopfield pueden almacenar informaci´on instant´aneamente sin necesidad de aprendiza je sin´aptico, y de una manera que puede ser todo lo robusto frente a ruido como se quiera (Johnson et al., 2011). Para ello el sistema aprovecha la existencia de estados metastables (situaciones que minimizan la energ´ıa del sistema localmente, sin corresponder al m´ınimo global) y como consecuencia aparecen transitorios en la dinmica de la actividad neuronal cuyas propiedades son consecuencia inmediata de las caracter´ısticas de la topolog´ıa subyacente y que es del tipo de las descritas anteriormente en el Cap´ıtulo 3 y en experimen- tos, esto es, el grado de agrupamiento o la modularidad de la red. B´asicamente, grupitos densamente interconectados de neuronas pueden mantener un estado conjunto de alta o ba ja actividad, en promedio. Considerando cada grupito como un elemento funcional elemental, en vez de cada neurona, se consigue la aparici´on de las propiedades requeridas. Es m´as, algunas otras caracter´ısticas 5 de la memoria de corto plazo emergen de manera natural de este mecanismo. En particular, se demuestra que la informaci´on se pierde gradualmente con el tiempo seg´un una ley aproximadamente potencial, como ha sido descrito en experimentos psicof´ısicos. En conclusi´on, las principales aportaciones originales de esta Tesis son: • M´etodos anal´ıticos y computacionales para estudiar redes evolutivas (Johnson et al., 2009b, 2010a) y redes con correlaciones nodo-nodo (Johnson et al., 2010b; de Franciscis et al., 2011). • Una respuesta a la pregunta de por qu´e la mayor´ıa de las redes reales son disasortativas (Johnson et al., 2010b). • Propiedades topol´ogicas que pueden optimizar el rendimiento de redes neuronales (Johnson et al., 2008; de Franciscis et al., 2011). • Un mecanismo que pudiera estar detr´as de la memoria de corto plazo (Johnson et al., 2011). Chapter 2 Where we are and where we’d like to go 2.1 From bridges to brains Strolling through the streets of Konigsberg, a young Immanuel Kant may have wondered whether, as some hoped, a path could be found that would take him once and only once over each of the city’s celebrated seven bridges and back to where he started. In 1736 Leonard Euler pointed out that for this or any other problem of the kind all that mattered was which land masses were connected to each other, and by how many bridges. In other words, the situation could be captured by a graph, as in Fig. 2.1, in which each land mass is represented by a node (also called vertex) and each bridge by a link (or edge). He showed that in the case of Konigsberg no such walk could be found, since an “Eulerian cycle” in a connected graph exists if and only if the degrees of all nodes are even numbers – the degree of a node being its number of edges (Euler, 1736). And thus was Graph Theory born. For over two centuries, the graphs people were interested in were precisely defined ob jects, usually sufficiently small to be drawn on a piece of paper. But in the late nineteen fifties, mathematicians began to study random graphs – i.e., defined only by some random generation process – perhaps with a view to better dealing with ever-growing communications networks (Bollob´as, 2001). E. N. Gilbert considered a situation in which there are n nodes and each pair is connected by an edge with probability q (Gilbert, 1959). For different values of these parameters, he was able to obtain the likelihood of the graph being connected (that is, of there being a path joining any two nodes). A similar model was proposed by Paul Erdos and Alfr´ed R´enyi: each of all the possi- 7 8 Chapter 2. Where we are and where we’d like to go Figure 2.1: The problem of the Seven Bridges of Konigsberg can be reduced to a graph in which nodes and edges represent land masses and bridges, re- spectively. ble graphs with n nodes and m edges had an equal chance of being “picked” (Erdos and R´enyi, 1959). In fact, a given graph will be generated with equal probability in either scenario, so the descriptions are equivalent, and usually known as the Erdos-R´enyi (ER) model. It is simple to see that if one were to average over many graphs generated by either of these processes, the degrees would follow a binomial distribution – tending, for large n, to a Poisson dis- tribution. That is, p(k) is symmetrically centred around its mean value and drops off exponentially – where ki is the degree of node i. An interesting phe- nomenon that can be observed using the ER model is that of percolation. If we measure the size Φ of the largest connected component (that is, of the highest number of nodes in the graph forming a connected subgraph) we obtain at different values of the probability q (or, equivalently, of m = 1 2 qn2), we see that there is a critical value, qc = 1/n, above which Φ/n does not vanish for high n – that is, there will usually exist a connected subgraph of a size compa- rable to that of the whole system. This passing from one situation (or phase) to another, each characterised by some qualitatively different characteristic, is known in physics as a phase transition. In this case, it is a second-order tran- sition, since the control parameter Φ varies continuously (and not abruptly, as in first-order transitions), and has innumerable applications. For instance, the nodes might be people susceptible to some disease, trees which may be set on fire, or oil bubbles in a porous medium. The epidemic will spread, the forest will burn, or the oil will be extractable if the density of edges – contagious contact, fire-conducting proximity, or links between pores – is over the critical 2.1 From bridges to brains 9 value. In his 1929 short story Chains (L´ancszemek, in the original Hungarian) Frigyes Karinthy suggested that the number of people in a chain of acquain- tances grows exponentially with size, and thus that very few steps are needed to join anyone with any other person. This Small World idea was taken up in 1967 by Stanley Milgram, who performed a series of experiments that, while somewhat less controversial than his well-known obedience-to-authority explo- rations, have nonetheless been widely discussed (Milgram, 1967). He and his colleagues sent various letters to random people with the request to attempt to send them on to a particular individual many thousands of miles away, plus the constraint that this had to be done via people with whom the sender was on first-name terms. Many of the letters reached their destination, after having been sent on by a surprisingly small number of intervening people. This was later popularised as the Six Degrees of Separation – the famous idea that any two people are linked by a path of only six acquaintances. Within the con- nected component of an ER random graph any two nodes are also joined by a short path – of the order of the logarithm of the number of nodes. However, this is less surprising, since these networks are not clustered; that is, they do not have the property typical of social networks whereby “the friends of my friends are (also likely to be) my friends.” In 1998, Duncan Watts and Steven Strogatz put forward a network model which took this feature into account. They considered a ring of n nodes, each connected to their k nearest neigh- bours (they set k = 4). Each edge was then broken from one of its nodes and rewired to some other random node with a probability p. Thus, p = 0 leaves the ring intact, while p = 1 changes it into an ER random graph. Two magni- tudes were measured for different values of p: the mean-minimum-path length, l, and the clustering coefficient, C . The first is simply an average over all pairs of nodes of the minimum-paths connecting them. The latter can be seen as the probability that two neighbours of a given node are directly connected to each other. For p = 0, the clustering is high (C = 1 2 ) and independent of the network size, while the mean minimum path scales with n (l ≃ n/8). At the other extreme, p = 1 yields a vanishingly small clustering (C = k/n) but short paths (l ≃ ln n/ ln k). The most interesting case is found at intermediate values of p. As p grows from zero, l falls very rapidly to a value close to the random case, but C does not present this drop until a much higher value is reached. Watts and Strogatz called this intermediate zone the Small-World region, since everyone is highly-interconnected while much of the local struc- 10 Chapter 2. Where we are and where we’d like to go ture is conserved. They suggested that this is actually a property of many real networks (as has since turned out to be the case (Newman, 2003c)), most especially of social networks – in which C is often several orders of magnitude greater than if the graph were random, while l is not much larger than in such a case. As the authors point out, it is essential to take this feature into account for the study of, say, epidemics. Another feature of networks which is quite ubiquitous in the real world is that degree distributions are highly heterogeneous; in fact, they often follow power-laws, p(k) ∼ k−γ , with γ a positive constant typically between 2 and 3. Such networks are nowadays referred to as scale free. In the nineteen fifties, Herbet Simon showed that these distributions come about when “the rich get richer” (Simon, 1955). Applying this idea to the case of scientific citations, Derek de Solla Price proposed the first known model of a scale-free network, in which nodes represent papers and edges are citations (de S. Price, 1965). Each node has an in-degree (the number of papers citing it) and an out-degree (papers it cites). That is, the network is directed, since edges have a direc- tion. Assuming that the probability a paper has of being cited by a new one is proportional to the number already citing it (its in-degree), the network is built up through the gradual addition of nodes, the neighbours of these be- ing chosen according to their existing in-degrees. Price showed analytically that such a mechanism leads to an in-degree distribution p(k) ∼ k−(2+1/m) , with m a parameter of the model equivalent to the mean degree1 . He called this mechanism cumulative advantage. Somewhat ironically – considering that Price, with a PhD in history from Cambridge, is best known as the father of scientometrics – this work was mostly ignored by the scientific community. The model was rediscovered in 1999 by Albert-L´aszl´o Barab´asi and Reka Al- bert, with the difference that they considered the network to be undirected (Barab´asi and Albert, 1999). They coined the term preferential attachment for the rich-get-richer mechanism, which is now generally assumed to be be- hind the formation of most scale-free networks (although other mechanisms exist (Caldarelli et al., 2002; Krapivsky et al., 2000; Newman, 2005)). Among many interesting consequences of such degree heterogeneity, Mark Newman showed that the clustering and mean-minimum-path length are respectively higher and lower than in homogeneous networks, making all scale-free net- works to some extent small worlds (Newman, 2003b). It also has important 1Note that in a directed network, the mean in-degree and mean out-degree coincide. 2.1 From bridges to brains 11 consequences for dynamical processes taking place among elements on the net- work, such as the synchronisation of coupled oscillators (Barahona and Pecora, 2002). As mentioned above, networks can be made up of separate components such that no path exists between nodes in different subgraphs. This is an extreme case of community structure. However, what is usually more inter- esting is the fact that communities may exist such that there is a higher density of edges within them than without, even if the network is connected (Girvan and Newman, 2002). These communities are also at times called mod- ules or clusters (although this can create some confusion with the related but distinct idea of clustering referred to above). Given a network, one can make a partition – that is, divide the nodes up into groups – and calculate what proportion of the edges fall within these, compared with the random expecta- tion. This measure is called the modularity of this partition, and sometimes one speaks of the modularity of a graph referring to that of the partition for which this value is maximum. Determining the community structure of empirical networks can often provide useful insights into aspects of the sys- tems. For instance, the communities may correspond to functional groups in a metabolic network, or groups of people who share some trait. How- ever, there are many problems related to making this kind of measurement. For one thing, there are so many possible partitions that even an ER ran- dom graph can have a fairly high modularity due simply to statistical fluctua- tions (Reichardt and Bornholdt, 2006). Then there is the fact that community structure can exist on may different levels – that is, the groups considered can be of any size – so one must usually consider a hierarchy of modules (Arenas et al., 2006). Furthermore, finding an optimal partition is an NP- Complete problem (Garey and Johnson, 1979), which makes comparing the modularity of each possible partition intractable in all but very small net- works. For these and other reasons, in recent years much work has gone into finding efficient algorithms to determine the community structure of networks, albeit approximately (Girvan and Newman, 2002; Donetti and Munoz, 2004; Blondel et al., 2008) – as well as into comparing the results offered by each approach (L. Danon et al., 2005). Finally, another feature of networks worth mentioning arises when the nodes of a network are endowed with some property and this is reflected by the layout of the edges: the situation is called a mixing pattern (Newman, 2002, 2003a). For instance, people might tend to choose sexual partners who 12 Chapter 2. Where we are and where we’d like to go share certain characteristics, such as mother-tongue or self-defined race. In these cases the network is assortative, since nodes of a kind assort, or group together. However, if the property in question were, say, gender, then the same graphs would be disassortative if most of the relations were heterosexual. In these cases the property can be considered discrete, but it can be continuous – for instance, people might assort according to age. An interesting case is when the property in question is the degree of each node, since it is then an entirely topological issue. The extent to which the degrees of neighbouring nodes are correlated – as given, for example, by Pearson’s correlation coeffi- cient (Newman, 2002) – is then a measure of the assortativity of the graph, being positive for assortative networks and negative for disassortative ones. It turns out that there is a striking universality in the nature of the degree- degree correlations displayed by real-world networks, whether natural or artifi- cial: social networks, like the ones just described, tend to be assortative, while almost all other kinds of network are disassortative (Pastor-Satorras et al., 2001; Newman, 2003c). Often specific functional constraints can be found to justify correlations of one or other kind, but in Chapter 5 of this thesis the purely topological explanation put forward in Ref. (Johnson et al., 2010b) is described. In any case, the degree-degree correlations of a network can play a significant role in the behaviour of processes taking place thereon. For exam- ple, assortative networks have lower percolation thresholds and are more robust to targeted attack (Newman, 2003a), while disassortative ones make for more stable ecosystems and seem to be more synchronizable (Brede and Sinha). All the aspects of networks mentioned in this brief overview, as well as many others, have been shown to be relevant for a wide range of complex systems (Albert and Barab´asi, 2002; Newman, 2003c; Boccaletti et al., 2006). Among these is the brain, a paradigm of complexity as well as the least understood of our organs. However, research focusing on how the collective behaviour of neural systems, as observed in mathematical models, is influenced by the topology of the underlying network is relatively scarce. This is perhaps due in part to the attention that other biological properties of the nervous system have tended to draw from the Computational Neuroscience community. Thanks to the flurry of activity that the field of Complex Networks has been enjoying over the last decade, this is a particularly good moment to undertake a more systematic analysis of how dynamics and topology are related in this kind of systems. 2.2 Neural networks in neuroscience 13 2.2 Neural networks in neuroscience Ever since the publication of Santiago Ram´on y Ca jal’s drawings of neurons – in his words, those “mysterious butterflies of the soul” – it has been clear that the nervous system is composed of a large number of such cells connected to one another to form a network (y Ca jal, 1995). Long axons, ending in termi- nals which form synapses to the dendrites which branch out from neighbouring neurons, transmit action potentials (APs) – changes in the cellular membrane voltage – and enable neurons somehow to cooperate and give rise the aston- ishing emergent phenomenon called thought. One of these APs is formed each time the membrane electric potential of a neuron surpasses a threshold value, leading to the opening of a great many voltage-dependent ionic gates between the cell and the extra-cellular medium. In turn, the membrane potential of a given neuron is constantly affected by action potentials arriving from neigh- bouring neurons, and thus an extremely complex web of cellular signalling is achieved. The first model neuron was proposed by McCulloch and Pitts (1943). This was simply an element that would return a Heaviside step function of the sum of its inputs. Sets of such “artificial neurons” could be used to implement any logical gate. Shortly after this, another important suggestion was made, this time by the psychologist Donald Hebb. Attempting to relate Pavlovian conditioning experiments with cellular plasticity, he conjectured, in 1949, the existence of some biological mechanism that would lead to neurons which re- peatedly fired (i.e., let off action potentials) together becoming more strongly coupled (Hebb, 1949). The initiation and propagation of action potentials in individual neurons was first modelled mathematically by Alan L. Hodgkin and Andrew Huxley in 1952 by means of a set of nonlinear ordinary differential equations which took into account the various ion fluxes (Hodgkin and Huxley, 1952). However, the concept of a neural network (as understood in theoretical and computational neuroscience) was partly inspired by mathematical models of spin systems. The first of these was the Ising model, put forward in 1920 by Wilhelm Lenz and studied by Ernst Ising with a view with a view to un- derstanding phase transitions and magnets (Onsager, 1944; Brush, 1967). It was known that the spin of electrons conferred a magnetic moment to indi- vidual atoms, but it wasn’t clear how exactly a very many such spins could self-organise into a large body producing a net magnetic field. By considering an infinite set of entities (spins) with possible values plus or minus one (up 14 Chapter 2. Where we are and where we’d like to go Figure 2.2: Drawing of the cells of the chick cerebellum, from “Estructura de los centros nerviosos de las aves”, Madrid, 1905. Notice how the neurons make up a complex network of synaptic interactions. or down, say) which, when placed at the nodes of a lattice, interact in such a way that energy is lowest when neighbours are aligned, and a temperature pa- rameter to govern the extent of random fluctuations, it was eventually shown that, below a certain critical temperature (in two or more dimensions), sym- metry is spontaneously broken and most of the spins end up aligned (Baxter, 1982). This ferromagnetic solution comes about and is then maintained be- cause it has a lower energy than any other configuration of spins. Subsequent models, in particular that of Sherrington and Kirkpatrick (1975), incorporated inhomogeneities in the coupling strengths such that there was no longer a con- figuration which simultaneously minimized all interaction energies, leading to frustrated states (spin-glasses). These ideas were put together, by Amari (1972) and then by Hopfield (1982), in the first neural network models to exhibit the mechanism known as associative memory. Each model neuron was placed at the node of a network, originally assumed to be fully connected (all nodes connected to all the rest), and followed a dynamics which can be seen either as that of Ising spins or of McCulloch-Pitts neurons. However, a noise parameter usually referred to as “temperature” by analogy with spin systems could be included to allow for non-deterministic behaviour. By setting the interaction strengths (synaptic weights) not randomly, as in the Sherrington-Kirkpatrick model, but according 2.2 Neural networks in neuroscience 15 to the Hebb rule referred to above, information could be stored and retrieved by the system. More specifically, a set of particular patterns, or configurations of positive and negative elements (firing and non-firing neurons), are recorded in the following way: for each pattern, one looks at each pair of neurons and adds a quantity to the weight of the synapse joining them if the pattern in question requires them to be in the same state, and subtracts it when they should be opposite. In this way, the minimum energy configurations correspond to the stored patterns, which therefore become attractors of the dynamics: if the temperature is not too heigh to destroy all order, the system will evolve towards whichever of these patterns most resembles the initial configuration it is placed in. Figure 2.3 illustrates how this mechanism works for a system such that the firing and non-firing neurons represent black and white pixels of a bitmap image. Thanks to associative memory, if we were to store, say, a set of photos of various people and then “show” the network a different picture of one of the same sub jects, it would be able to retrieve the correct identity. Not only is this mechanism used nowadays in technology capable of performing tasks such as pattern discrimination and classification, but it is widely considered to underlie our own capacity for learning and recalling information. There is evidence from neuronal readouts that this is so (Amit, 1995), and not long ago, in vivo experiments finally established that learning is indeed related to the processes of long term potentiation (LTP) and depression (LTD) – by which synapses between neurons that fire nearly simultaneously gradually increase or decrease their conductance depending on the interval of time elapsed (Gruart et al., 2006; Roo et al., 2008). The neural network models studied nowadays generally include more real- istic dynamics both for the neurons and for the synapses, taking into account a variety of cellular and subcellular processes (Amit, 1989; Torres and Varona, 2010). For example, the fact that the conductance of synapses in reality de- pends on their workload has been found to enable a network to switch from one pattern to another – either spontaneously or as a reaction to sensory stim- uli – providing a means for the performance of dynamic tasks (Cortes et al., 2006; Holcman and Tsodyks, 2006); this result also seems to agree well with physiological data (Korn and Faure, 2003). In fact, there is evidence that the brain somehow maintains itself close to a boundary – called, in physics, a critical point – between an ordered and a chaotic regime (Egu´ıluz et al., 2005; Chialvo, 2004; Chialvo et al., 2008; Bonachela et al., 2010; Torres and Varona, 16 Chapter 2. Where we are and where we’d like to go Figure 2.3: In the Hopfield neural-network model, the interaction strengths (representing synaptic weights) store information in the form of particular patterns, or configurations of firing neurons, which become attractors of the dynamics. Whatever the initial state of the system, it will always evolve to- wards one of these patterns, thus allowing for the storage and retrieval of information. The mechanism, known as associative memory, is thought to be at the basis of memory in the brain. In this case, the network is “remember- ing” an illustration by Jean-Baptiste Oudry for Jean de la Fontaine’s fable La Cigale et la Fourmi. 2010). This would be in line with research that shows how certain useful prop- erties – such as the computational capacity of some neural-network models (Bertschinger and Natschlager, 2004), or the dynamic range of sensitivity to stimuli in sensory systems (Kinouchi and Copelli, 2006) – are optimised at this “edge of chaos (Chialvo, 2006)”. That these models should actually reflect, albeit in an enormously simpli- fied way, what actually goes on in our brains tends to fit in quite well with intuitive expectations – to the extent that so-called connectionist models seem to be gradually becoming the accepted paradigm in relevant areas of psychol- ogy and philosophy (Marcus and G.F., 2001; Frank, 1997). For instance, from 2.2 Neural networks in neuroscience 17 this point of view the way in which the recollection of a particular detail often evokes, almost instantly, a whole landscape of sensations and emotions makes sense, since these concepts will have been stored in some way as the same pat- tern. Also, the fact that new memories are recorded in synapses which were already being used to store previous information would seem to explain why memories tend to fade slowly with time, yet can still be recalled, at least to some extent, when a particular thought in some sense overlaps with (reminds one of ) one of them. When this happens, the old memory springs to mind and, if held there for long enough, can be refreshed via long term potentiation and depression – although interaction with other patterns or with current stimuli may well modify the refreshed information. Similarly, previous information influences the storing of new memories, leading to the well known fact that we tend to “see” things the way we expect them to be. It seems, then, that the basic mechanisms behind the ability of our brains to remember things, at least when the information is stored slowly enough for the biochemical processes of LTP and LTD to be at work (long-term memory), are now understood. Not only are the implications of such knowledge far- reaching in themselves. The way in which it was developed is also particularly notworthy. More or less sketchy ideas from areas as diverse as behavioural psychology, neurophysiology and theoretical physics were brought together in order to come up with a minimal mathematical model capable of manifesting the sought-after phenomenon of information retrieval as a consequence of the known properties of a great many simple elements. This kind of research can at first seem more like a mathematical game than anything to do with nitty- gritty reality. But the fact the basic mechanism of associative memory has since borne up to decades of experimenting and theoretical probing reveals how insightful it can actually be. It is likely that other features of brain function – short-term memory, information processing or emotional tagging, to name but the first few that spring to mind – will eventually be thrown under a similar light. In fact, we can expect the nature of even such an elusive and intimate phenomenon as consciousness some day to become clear. After all, the explanations behind other emergent properties of matter which in their day seemed almost mystical, such as temperature or life, are now well understood. 18 Chapter 2. Where we are and where we’d like to go 2.3 A declaration of intent As Zora Neale Hurston put it, “Research is formalized curiosity. It is poking and prying with a purpose.” But there are many possible purposes, and even more different ways of poking and prying. The motivation behind the work presented here is to understand how the phenomena we observe in certain systems on a macroscopic scale can come about from interactions of their many relatively simple constituent elements. In the case of neural systems, it seems reasonable to assume that these basic elements are neurons, and that it is thanks to the cooperation of a great many of these cells that such organs are able to think and feel. The human brain – with about 100 billion neurons connected by 100 trillion synapses – being the most complex system we know of, an enormous degree of simplification will be required for our description to be of any use to this purpose. (In fact, if we could somehow simulate a brain in all detail, the result would be just as unfathomable as the original ob ject, however exciting the activity may prove for other reasons.) The physiology of the neuron is nowadays quite well understood. However, just as the properties of atoms or transistors that are key to understanding phase transitions or the workings of a microchip are, respectively, magnetic interactions and voltage- dependent gating, we must try to ascertain exactly which neuronal features are necessary for the macroscopic behaviour we are interested in to occur. One way to do this is to start by considering only the most basic characteristics and explore what non-trivial phenomena emerge from these, allowing us then to add new ingredients one at a time to pinpoint the relevant ones. In this line, we consider large sets of Hopfield’s simple binary model neurons to study how network properties are related to collective behaviour. This work is laid out as follows. Chapter 3 deals with development. The appearance and disappearance of edges in a network (growth and death of synapses, in the case of the brain) is formalised as a stochastic process and studied in a general setting (Johnson et al., 2009b, 2010a). It turns out that many of the topological features observed in experiments are well modelled in this way – which to some extent justifies, a posteriori, our initial assumptions. The following chapters describe particular phenomena that emerge as a di- rect consequence of some of those topological features: degree heterogeneity in conjunction with synaptic depression improves the performance of dynamical tasks (Johnson et al., 2008) (Chapter 4); assortativity serves to enhance a neu- ral network’s robustness to noise (de Franciscis et al., 2011) (Chapter 6); and clustering or modularity can lead to metastable states with certain properties 2.3 A declaration of intent 19 essential for some short-term memory abilities (properties hitherto lacking in previous models) (Johnson et al., 2011) (Chapter 7). Thanks to the extreme simplicity of the basic elements we are considering, we are able not only to simulate but also to understand mathematically how exactly the interesting phenomena emerge. This makes it possible to predict, to some extent, which extra ingredients wil l not invalidate the results if they are taken into account explicitly. Some of the work has a more general scope than the study of neural net- works. In particular, the equations obtained in Chapter 3 can be applied to any network that evolves under the influence of probabilistic addition and deletion of edges. And the method put forward in Chapter 5 for the study of correlated networks can be used not just for analysing particular models, as we go on to do in Chapter 6, but to solve many other problems – such as that of the ubiquity of disassortative networks in nature and technology (Johnson et al., 2010b), or how the property of nestedness typical of ecosystems is related to other topological characteristics (c.f. Appendix C). To sum up, the aim of the thesis is to shed light on how cellular dy- namics can lead to the complex network structures of neural systems, and, in its turn, in what ways this topology can influence, optimise and determine the collective behaviour of such systems. The main contributions made are: • An analytical method to study the evolution of networks governed by a combination of local and global stochastic rules. • A mathematical and computational technique for the study of correlated networks in a model-independent way. • Possible biological justifications for two non-trivial features of the topol- ogy of the human cortex: heterogeneity of the degree distribution and high assortativity. • An answer to the long-standing question of why most networks are dis- assortative. • Cluster Reverberation: the first mechanism proposed which would allow neural systems to store information instantaneously in a robust manner. Chapter 3 Evolving networks and the development of neural systems The highly heterogeneous degree distributions of most empirical networks is assumed in many cases to arise from some form of cummulative advantage, or preferential attachment. However, the origin of various other topological features is often not clear and attributed to specific functional requirements. We show how it is possible to analyse a very general scenario in which nodes gain or lose edges according to arbitrary functions of local and/or global degree information. Applying our method to two rather different examples of brain development – synaptic pruning in humans and the neural network of the worm C. Elegans – we find that simple biologically motivated assumptions lead to very good agreement with experimental data. In particular, many nontrivial topological features of the worm’s brain arise naturally at a critical point. 3.1 Introduction The conceptual simplicity of a network – a set of nodes, some pairs of which connected by edges – often suffices to capture the essence of cooperation in complex systems. Ever since Barab´asi and Albert presented their evolving network model (Barab´asi and Albert, 1999), in which linear preferential at- tachment leads asymptotically to a scale-free degree distribution (the degree, k , of a node being its number of neighbouring nodes), there have been many variations or refinements to the original scenario (Albert and Barab´asi, 2000; G. Bianconi and Barab´asi, 2001; Krapivsky et al., 2000; Bianconi and Barab´asi, 2001; Park et al., 2005; Ree, 2007) (see also the review by Boccaletti et al. (2006)). In Ref. (Johnson et al., 2009b), we show how topological phase tran- 21 Chapter 3. Evolving networks and the development of neural systems 22 sitions and scale-free solutions can emerge in the case of nonlinear rewiring in fixed-size networks, and this work is summarized in Appendix A. In Ref. (Johnson et al., 2010a) we extend our scope to more general and realistic situa- tions, considering the evolution of networks making only minimal assumptions about the attachment/detachment rules. In fact, all we assume is that these probabilities factorize into two parts: a local term that depends on node de- gree, and a global term, which is a function of the mean degree of the network. This is the work described in this chapter. Our motivation can be found in the mechanisms behind many real-world networks, but we focus, for the sake of illustration, on the development of bio- logical neural networks, where nodes represent neurons and edges play the part of synaptic interaction (Amit, 1989; Sporns et al., 2004; Torres and Varona, 2010). Experimental neuroscience has shown that enhanced electric activity induces synaptic growth and dendritic arborization (Lee et al., 1980; Frank, 1997; Klintsova and Greenough, 1999; Roo et al., 2008). Since the activity of a neuron depends on the net current received from its neighbours, which tends to be higher the more neighbours it has, we can see node degree as a proxy for this activity – accounting for the local term alluded to above. On the other hand, synaptic growth and death also depend on concentrations of various molecules, which can diffuse through large areas of tissue and there- fore cannot in general be considered local. A feature of brain development in many animals is synaptic pruning – the large reduction in synaptic den- sity undergone throughout infancy. Chechik et al. (1999, in press) have shown that via an elimination of less needed synapses, this can reduce the energy consumed by the brain (which in a human at rest can account for a quarter of total energy used) while maintaining near optimal memory performance. Going on this, we will take the mean degree of the network – or mean synaptic density – to reflect total energy consumption, hence the global terms in our attachment/detachment rules (Johnson et al., 2009a). An alternative approach would be to consider some kind of model neu- rons explicitly and couple the probabilities of synaptic growth and death to neuronal dynamic variables, such as local and global fields. In an Amari- Hopfield network, for example, the expected value of the field (total incoming current) at node i is proportional to i’s degree (Torres et al., 2004), the to- tal current (energy consumption) in the network therefore being proportional to the mean degree; qualitatively, these observations are likely to hold also in more realistic situations (Magistretti, 2009), although relations need not 3.2 Basic considerations 23 be linear. Co-evolving networks of this sort are currently attracting atten- tion, with dynamics such as Prisoner’s Dilemma (Poncela et al., 2008), Voter Model (Vazquez et al., 2008) or Random Walkers (Antiqueira et al., 2009). Although we consider this line of work particularly interesting, for generality and analytical tractability we opt here to use only topological information for the attachment/detachment rules, although our results can be applied to any situation in which the dynamical states of the elements at the nodes can be functionally related to degrees1 . Following a brief general analysis, we show how appropriate choices of func- tions induce the system to evolve towards heterogeneous (sometimes scale-free) networks while undergoing synaptic pruning in quantitative agreement with experiments. At the same time, degree-degree correlations emerge naturally, thus making the resulting networks disassortative – as tends to be the case for most biological networks – and leading to realistic small-world parameters. 3.2 Basic considerations Consider a simple undirected network with N nodes defined by the adjacency matrix a, the element aij representing the existence or otherwise of an edge between nodes i and j . Each node can be characterized by its degree, ki = Pj aij . Initially, the degrees follow some distribution p(k , t = 0) with mean κ(t). We wish to study the evolution of networks in which nodes can gain or lose edges according to stochastic rules which only take into account local and global information on degrees. So as to implement this in the most general way, we will assume that at every time step, each node has a probability of gain gaining a new edge, P , to a random node; and a probability of losing a i gain randomly chosen edge, P lose . We assume these factorize as P = u(κ)π(ki) i i and P lose = d(κ)σ(ki), where u, d, π and σ can be arbitrary functions, but i impose nothing else other than normalization. For each edge that is withdrawn from the network, two nodes decrease in degree: i, chosen according to σ(ki ), and j , a random neighbour of i’s; so there is an added effective probability of loss kj /(κN ). Similarly, for each edge placed in the network, not only l chosen according to π(kl ) increases its degree; 1For instance, the stationary distribution of walkers used for edge dynamics by Antiqueira et al. (2009) is actually obtained purely from topological information, although it can only be written in terms of local degrees for undirected networks. Chapter 3. Evolving networks and the development of neural systems 24 a random node m will also gain, with the consequent effective probability N −1 (though see2 ). Let us introduce the notation π(k) ≡ π(k) + N −1 and σ(k) ≡ σ(k) + k/(κN ). Network evolution can now be seen as a one step process (van Kampen, 1992) with transition rates u(κ) π(k) and d(κ) σ(k). The expected value for the increment in a given p(k , t) at each time step (which we equate with a temporal derivative) defines a master equation for the degree distribution (Johnson et al., 2009b): dp(k , t) dt = u (κ) π(k − 1)p(k − 1) + d (κ) σ(k + 1)p(k + 1) (3.1) (3.2) ∂pst (k) ∂k − [u (κ) π(k) + d (κ) σ(k)] p(k , t). Assuming now that p(k , t) evolves towards a stationary distribution, pst (k), then this must necessarily satisfy detailed balance since it is a one step process (van Kampen, 1992); i.e., the flux of probability from k to k + 1 must equal the flux from k + 1 to k , for all k (Marro and Dickman). This condition (sufficient for Eq. (3.1) to be zero) can be written as σ(k + 1) − 1(cid:21) pst(k), = (cid:20) u(κst) π(k) d(κst) where we have substituted a difference for a partial derivative and κst ≡ Pk kpst(k). Setting π and σ so as to be normalized to one (i.e., Pk p(k)π(k) = Pk p(k)σ(k) = 1, ∀t), which is equivalent to saying that at each time step ex- actly u(κ) nodes are chosen to gain edges and d(κ) to lose them, then in the stationary state we will have u(κst) = d(κst) since the total number of edges will be conserved. From Eq. (3.2) we can see that pst (k) will have an extremum at some value ke if it satisfies π(ke ) = σ(ke + 1). ke will be a maximum (mini- mum) if the numerator in Eq. (3.2) is smaller (greater) than the denominator for k > ke , and viceversa for k < ke . Assuming, for example, that there is one and only one such ke , then a maximum implies a relatively homogeneous distribution, while a minimum means pst(k) will be split in two, and therefore highly heterogeneous. More intuitively, if for nodes with large enough k there is a higher probability of gaining edges than of losing them, the degrees of these nodes will grow indefinitely, leading to heterogeneity. If, on the other hand, highly connected nodes always lose more edges than they gain, we will 2We are ignoring the small corrections that arise because j 6= i and l 6= m, which in any case would disappear if self-connections were allowed. 3.3 Synaptic pruning 25 obtain quite homogeneous networks. From this reasoning we can see that a particularly interesting case (which turns out to be critical) is that in which π(k) and σ(k) are such that π(k) = σ(k) ≡ v (k), ∀k . According to Eq. (3.2), Condition (3.3) means that for large k , ∂pst (k)/∂k → 0, and pst(k) flattens out – as for example a power-law does. The standard Fokker-Planck approximation for the one step process defined by Eq. (3.1) is (van Kampen, 1992): (3.3) ∂p(k , t) ∂ t = 1 2 ∂ 2 ∂k2 {[d(κ) σ(k) + u(κ) π(k)] p(k , t)} (3.4) ∂ ∂k {[d(κ) σ(k) − u(κ) π(k)] p(k , t)} . For transition rates which meet Condition (3.3), Eq. (3.4) can be written as: + ∂p(k , t) ∂ t = 1 2 [d (κ) + u (κ)] ∂ 2 ∂k2 [v (k)p(k , t)] (3.5) ∂ ∂k [v (k)p(k , t)] . + [d (κ) − u (κ)] Ignoring boundary conditions, the stationary solution must satisfy, on the one hand, v (k)pst(k) = Ak + B , so that the diffusion is stationary, and, on the other, u(κst) = d(κst), to cancel out the drift. For this situation to be reach- able from any initial condition, u(κ) and d(κ) must be monotonous functions, decreasing and increasing respectively. 3.3 Synaptic pruning As a simple example, we will first consider global probabilities which have the linear forms: and d[κ(t)] = , u[κ(t)] = κmax (cid:19) N (cid:18)1 − n κ(t) where n is the expected value of the number of additions and deletions of edges per time step, and κmax is the maximum value the mean degree can have. This choice describes a situation in which the higher the density of synapses, the less likely new ones are to sprout and the more likely existing ones are to atrophy – a situation that might arise, for instance, in the presence of a finite quantity κ(t) κmax (3.6) n N Chapter 3. Evolving networks and the development of neural systems 26 h (3.7) of nutrients. Again taking into account that π and σ are normalized to one, gain − P lose we find that the expected increment in κ(t) is summing over P i i N (cid:20)1 − 2 κmax (cid:21) ∆κ(t) κ(t) n ∆t i = 2{u[κ(t)] − d[κ(t)]} = 2 (independently of the local probabilities). Therefore, the mean degree will increase or decrease exponentially with time, from κ(0) to 1 2 κmax . Assuming that the initial condition is, say, κ(0) = κmax , and expressing the solution in terms of the mean synaptic density – i.e., ρ(t) ≡ κ(t)N/(2V ), with V the total volume considered – we have ρ(t) = ρf (cid:16)1 + e−t/τp (cid:17) , where we have defined ρf ≡ ρ(t → ∞) and the time constant for pruning is τp = ρfN/n. This equation was fitted in Fig. 3.1 to experimental data on layers 1 and 2 of the human auditory cortex3 obtained during autopsies by Huttenlocher and Dabholkar (1997). It seems reasonable to assume that the initial overgrowth of synapses is due to the transient existence of some kind of growth factors. If we account for these by including a nonlinear, time-dependent term g (t) ≡ a exp(−t/τg ) in the probability of growth, i.e., u[κ(t), t] = (n/N )[1 − κ(t)/κmax + g (t)], leaving d[κ(t)] as before, we find that ρ(t) becomes − t−t0 ρ(t) = ρf (cid:20)1 + e−t/τp − (cid:16)1 + e−t0 /τp (cid:17) e τg (cid:21) , where t0 is the time at which synapses begin to form (t = 0 corresponds to the moment of conception) and τg is the time constant related to growth. The inset in Fig. 3.1 shows the best fit to the auditory cortex data. Since the contour conditions ρf and (for Eq. (3.8)) t0 are simply taken as the value of the last data point and the time of the first one, in each case, the time constants τp and τg are the only parameters needed for the fit. (3.8) 3.4 Phase transitions The drift-like evolution of the mean degree we have just illustrated with the example of synaptic pruning is independent of the local probabilities π(k) 3Data points for three particular days (smaller symbols) are omitted from the fit, since we believe these must be from sub jects with inherently lower synaptic density. 3.4 Phase transitions 27 Cortical layer 1 Cortical layer 2 80 80 80 80 ) ) ) ) t t t t ( ( ( ( ρ ρ ρ ρ 0 0 0 0 100 10000 1000 100 100 100 1000 1000 1000 10000 10000 10000 Conceptual age (days) Conceptual age (days) Conceptual age (days) Conceptual age (days) 3 3 3 3 3 3 3 3 m m m m m m m m µ µ µ µ µ µ µ µ / / / / / / / / s s s s s s s s e e e e e e e e s s s s s s s s p p p p p p p p a a a a a a a a n n n n n n n n y y y y y y y y S S S S S S S S : : : : : : : : ) ) ) ) ) ) ) ) t t t t t t t t ( ( ( ( ( ( ( ( ρ ρ ρ ρ ρ ρ ρ ρ 80 80 80 80 80 80 80 80 70 70 70 70 70 70 70 70 60 60 60 60 60 60 60 60 50 50 50 50 50 50 50 50 40 40 40 40 40 40 40 40 30 30 30 30 30 30 30 30 20 20 20 20 20 20 20 20 0 0 0 0 0 0 0 0 5000 5000 5000 5000 5000 5000 5000 5000 10000 15000 10000 10000 10000 10000 10000 10000 10000 15000 15000 15000 15000 15000 15000 15000 Conceptual age (days) Conceptual age (days) Conceptual age (days) Conceptual age (days) Conceptual age (days) Conceptual age (days) Conceptual age (days) Conceptual age (days) 20000 20000 20000 20000 20000 20000 20000 20000 25000 25000 25000 25000 25000 25000 25000 25000 Figure 3.1: Synaptic densities in layers 1 (red squares) and 2 (black cir- cles) of the human auditory cortex against time from conception. Data from Huttenlocher and Dabholkar (1997), obtained by directly counting synapses in tissues from autopsies. Lines follow best fits to Eq. (3.7), where the param- eters were: for layer 1, τp = 5041 days; and for layer 2, τp = 3898 days (for ρf we have used the last data pints: 30.7 and 40.8 synapses/µm3 , for layers 1 and 2 respectively). Data pertaining to the first year and to days 4700, 5000 7300, shown with smaller symbols, where omitted from the fit. Assuming the existence of transient growth factors, we can include the data points for the first year in the fit by using Eq. (3.8). This is done in the inset (where time is displayed logarithmically). The best fits were: for layer 1, τg = 151.0 and τp = 5221; and for layer 2, τg = 191.1 and τp = 4184, all in days (we have approximated t0 to the time of the first data points, 192 days). and σ(k). The effect of these is rather in the diffusive behaviour which can lead, as mentioned, either to homogeneous or to heterogeneous final states. A useful bounded order parameter to characterize these phases is therefore m ≡ exp (−σ 2/κ2 ) , where σ 2 = hk2i − κ2 is the variance of the degree dis- tribution (h·i ≡ N −1 Pi (·) represents an average over nodes). We will use mst ≡ limt→∞ m(t) to distinguish between the different phases, since mst = 1 for a regular network and mst → 0 for one following a highly heterogeneous distribution. Although there are particular choices of probabilities which lead to Eq. (3.5), these are not the only critical cases, since the transition from Chapter 3. Evolving networks and the development of neural systems 28 100 100 100 100 100 100 100 Critical Critical Critical Critical Critical Critical Critical ) ) ) ) ) ) ) t t t t t t t , , , , , , , k k k k k k k ( ( ( ( ( ( ( p p p p p p p 10-2 10-2 10-2 10-2 10-2 10-2 10-2 10-4 10-4 10-4 10-4 10-4 10-4 10-4 ) ) t t ( ( κ κ 20 20 10 10 0 0 1 1 k-2 k-2 k-2 k-2 k-2 k-2 k-2 ) ) ) ) ) ) ) ) ) t t t t t t t t t ( ( ( ( ( ( ( ( ( m m m m m m m m m 2.5 103 2.5 103 0 0 5 104 5 104 0 0 Time (MCS) Time (MCS) t=102 t=103 t=106 100 100 100 100 100 100 100 Supercritical Supercritical Supercritical Supercritical Supercritical Supercritical Supercritical 10-2 10-2 10-2 10-2 10-2 10-2 10-2 10-4 10-4 10-4 10-4 10-4 10-4 10-4 ) ) t t ( ( κ κ 20 20 10 10 0 0 1 1 ) ) t t ( ( m m 2.5 103 2.5 103 0 0 5 104 5 104 0 0 Time (MCS) Time (MCS) k-4 k-4 k-4 k-4 k-4 k-4 k-4 k-4 k-4 k-4 k-4 t=102 t=103 t=105 100 100 100 100 100 100 100 101 101 101 101 101 101 101 102 102 102 102 102 102 102 k k k k k k k 103 103 103 103 103 103 103 100 100 100 100 100 100 100 101 101 101 101 101 101 101 102 102 102 102 102 102 102 k k k k k k k 103 103 103 103 103 103 103 Figure 3.2: Evolution of the degree distributions of networks beginning as regular random graphs with κ(0) = 20 in the critical (top) and supercritical (bottom) regimes. Local probabilities are σ(k) = k/(hkiN ) in both cases, and π(k) = 2σ(k)−N −1 and π(k) = k3/2/(hk3/2 iN ) for the critical and supercritical ones, respectively. Global probabilities as in Eq. (3.6), with n = 10 and κmax = 20. Symbols in the main panels correspond to p(k , t) at different times as obtained from MC simulations. Lines result from numerical integration of Eq. (3.1). Insets show typical time series of κ and m. Light blue lines are from MC simulations and red lines are theoretical, given by Eq. (3.7) and Eq. (3.1), respectively. N = 1000. homogeneous to heterogeneous stationary states can come about also with functions which never meet Condition (3.3). Rather, this is a classic topolog- ical phase transition, the nature of which depends on the choice of functions (Park and Newman, 2004; Burda et al., 2004; Der´enyi et al., 2004) . Evolution of the degree distribution is shown in Fig. 3.2 for critical and supercritical choices for the probabilities, as given by MC simulations (starting from regular random graphs) and contrasted with theory. (The subcritical regime is not shown since the stationary state has a distribution similar to the ones at t = 103 MCS in the other regimes.) The disparity between the theory and the simulations for the final distributions is due to the build up of certain correlations not taken into account in our analysis. This is because the existence of some very highly connected nodes reduces the probability of there being very low degree nodes. In particular, if there are, say, x nodes connected to the rest of the network, then a natural cutoff, kmin = x, emerges. Note that this occurs only when we restrict ourselves to simple networks, i.e., with only one edge allowed for each pair of nodes. This topological phase transition is shown in Fig. 3.3, where mst is plotted against parameter α for global 3.4 Phase transitions 29 0 0 0 0 0 0 r r r r r r -1 -1 -1 -1 -1 -1 0.5 0.5 0.5 0.5 0.5 0.5 1.5 1.5 1.5 1.5 1.5 1.5 2 2 2 2 2 2 1 1 1 1 1 1 α α α α α α 1 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 0 0 t t t t t t t t t s s s s s s s s s m m m m m m m m m N=1000 N=1500 N=2000 103 103 Ν Ν λ λ Q Q λ λ N N 0 0 0.5 0.5 0.5 0.5 1.5 1.5 1.5 1.5 2 2 2 2 1 1 1 1 α α α α 5 102 5 102 Q Q 0 0 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 1 1 1 1 1 1 1 1 1 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 2 2 2 2 2 2 2 2 2 α α α α α α α α α Figure 3.3: Phase transitions in mst for π(k) ∼ kα and σ(k) ∼ k , and u(κ) and d(κ) as in Eq. (3.6). N = 1000 (blue squares), 1500 (red triangles) and 2000 (black circles); κ(0) = κmax = 2n = N/50. Corresponding lines are from numerical integration of Eq. (3.1). The bottom left inset shows values of the highest eigenvalue of the Laplacian matrix (red squares) and of Q = λN /λ2 (black circles), a measure of unsynchronizability; N = 1000. The top right inset shows transitions for the same parameters in the final values of Pearson’s correlation coefficient r (see Section 3.5), both for only one edge allowed per pair of nodes (red squares) and without this restriction (black circles). probabilities as in Eq. (3.6) and local ones π(k) ∼ kα and σ(k) ∼ k . This situation corresponds to one in which edges are eliminated randomly while nodes have a power-law probability of sprouting new ones (note that power- laws are good descriptions of a variety of monotonous response functions, yet only require one parameter). Although, to our knowledge, there are not yet enough empirical data to ascertain what degree distribution the structural topology of the human brain follows, it is worth noting that its functional topology, at the level of brain areas, has been found to be scale-free with an exponent very close to 2 (Egu´ıluz et al., 2005). In general, most other measures can be expected to undergo a transition along with its variance. For instance, highly heterogeneous networks (such as scale-free ones) exhibit the small-world property, characterized by a high clustering coefficient, C ≫ hki/N , and a low mean minimum path, l ∼ ln(N ) (Watts and Strogatz, 1998). A particularly interesting topological feature of Chapter 3. Evolving networks and the development of neural systems 30 a network is its synchronizability – i.e., given a set of oscillators placed at the nodes and coupled via the edges, how wide a range of coupling strengths will result in them all becoming synchronized. Barahona and Pecora showed analytically that, for linear oscillators, a network is more synchronizable the lower the relation Q = λN /λ2 – where λN and λ2 are the highest and lowest non-zero eigenvalues of the Laplacian matrix ( Λij ≡ δij ki − aij ), respectively (Barahona and Pecora, 2002). The bottom left inset in Fig. 3.3 displays the values of Q and λN obtained for the different stationary states. There is a peak in Q at the critical point. It has been argued that this tendency of heterogeneous topologies to be particularly unsynchronizable poses a paradox given the wide prevalence of scale-free networks in nature, a problem that has been deftly got around by considering appropriate weighting schemes for the edges (Motter et al., 2005; Chavez et al., 2005) (see also4 , and the review by Arenas et al. (2008a)). However, there is no generic reason why high synchro- nizability should always be desirable. In fact, it has recently been shown that heterogeneity can improve the dynamical performance of model neural net- works precisely because the fixed points are easily destabilised (Johnson et al., 2008) (as well as conferring robustness to thermal fluctuations and improving storage capacity (Torres et al., 2004)). This makes intuitive sense, since, pre- sumably, one would not usually want all the neurons in one’s brain to be doing exactly the same thing. Therefore, this point of maximum unsynchronizability at the phase transition may be a particularly advantageous one. On the whole, we find that three classes of network – homogeneous, scale- free (at the critical point) and ones composed of starlike structures, with a great many small-degree nodes connected to a few hubs – can emerge for any kind of attachment/detachment rules. It follows that a network sub ject to some sort of optimising mechanism, such as Natural Selection for the case of living systems, could thus evolve towards whichever topology best suits its requirements by tuning these microscopic actions. 3.5 Correlations Most real networks have been found to exhibit degree-degree correlations, also known as mixing by degree (Pastor-Satorras et al., 2001; Newman, 2003c). 4Using pacemaker nodes, scale-free networks have also been shown to emerge via rules which maximize synchrony (Sendina-Nadal et al., 2008). 3.5 Correlations 31 They can thus be classified as assortative, when the degree of a typical node is positively correlated with that of its neighbours, or disassortative, when the correlation is negative. This property has important implications for net- work characteristics such as connectedness and robustness (Newman, 2002, 2003a). A useful measure of this phenomenon is Pearson’s correlation coef- ficient applied to the edges (Newman, 2003c,a; Boccaletti et al., 2006): r = ([klk ′ l ] − [kl ]2)/([k2 l ] − [kl ]2), where kl and k ′ l are the degrees of each of the two nodes pertaining to edge l, and [·] ≡ (hkiN )−1 Pl (·) represents an average over edges; r ≤ 1. Writing Pl (·) = Pij aij (·), r can be expressed in terms of averages over nodes: r = hkihk2knn(k)i − hk2i2 (3.9) , hkihk3i − hk2i2 where knn(k) is the mean nearest-neighbour-degree function; i.e., if knn,i ≡ k−1 i Pj aij kj is the mean degree of the neighbours of node i, knn (k) is its average over all nodes such that ki = k . Whereas most social networks are assortative (r > 0) – due, probably, to mechanisms such as homophily (Newman, 2003c) – almost all other networks, whether biological, technological or information- related, seem to be generically disassortative. The top right inset in Fig. 3.3 displays the stationary value of r obtained in the same networks as in the main panel and lower inset. It turns out that the heterogeneous regime is disassorta- tive, the more so the larger α. (Note that a completely homogeneous network cannot have degree-degree correlations, since all degrees are the same.) It is known that the restriction of having at most one edge per pair of nodes induces disassortativity (Park and Newman, 2003; Maslov et al., 2004). However, in our case this is not the sole origin of the correlations, as can also be seen in the same inset of Fig. 3.3, where we have plotted r for networks in which we have lifted the restriction and allowed any number of edges per pair of nodes. In fact, when multiple edges are allowed, the correlations are slightly stronger. As we shall discuss in Chapter 5, there is a general entropic reason for hetero- geneous networks, in their equilibrium state (i.e., in the absence of correlating mechanisms), to become disassortative (Johnson et al., 2010b). But neither is this here the case, since the networks generated are driven from equilibrium by the mechanisms of preferential attachment and detachment. To understand how these specific correlations come about, consider a pair of nodes (i, j ), which, at a given moment, can either be occupied by an edge or unoccupied. We will call the expected times of permanence for occupied and unoccupied states τ o ij and τ u ij , respectively. After sufficient evolution time (so Chapter 3. Evolving networks and the development of neural systems 32 that occupancy becomes independent of the initial state5 ), the expected value of the corresponding element of the adjacency matrix, E (aij ) ≡ ǫij , will be τ o ij ij + τ u τ o ij ǫij = . If p+ ij (p− ij ) is the probability that (i, j ) will become occupied (unoccupied) given ij and τ u that it is unoccupied (occupied), then τ o ij ∼ 1/p+ ij ∼ 1/p− ij , yielding ij !−1 ǫij = 1 + p− ij p+ Taking into account the probability that each node has of gaining or losing an edge, we obtain6 : p+ ij = u(hki)N −1 [π(ki ) + π(kj )] and p− ij = d(hki)[σ(ki)/ki + ij ≫ p+ σ(kj )/kj ]. Then, assuming that the network is sparse enough that p− ij (since the number of edges is much smaller than the number of pairs), and particularising for power-law local probabilities π(k) ∼ kα and σ(k) ∼ kβ , the expected occupancy of the pair is . = p+ ij p− ij ǫij ≃ hkαiN kα j ! . i + kα hkβ i u(hki) j kβ−1 i + kβ−1 d(hki) Considering the stationary state, when u(hki) = d(hki), and for the case of random deletion of edges, β = 1 (so that the only nonlinearity is due to α), the previous expression reduces to ǫij ≃ hki 2hkαiN (cid:0)kα i + kα j (cid:1) . (Note that this matrix is not consistent term by term, since Pj ǫij 6= ki , although it is globally consistent: Pij ǫij = hkiN .) The nearest-neighbour- degree function is now 1 ǫij kj = hki ki Xj i + hkα+1 ik−1 (hkikα−1 knn (ki ) = i ) 2hkαi (a decreasing function for any α), with the result that Pearson’s coefficient becomes hkα i (cid:18) hki3hkα+1 i − hk2i2 hkαi (cid:19) . 1 hkihk3i − hk2i2 5Note that this will always happen eventually since the process is ergodic. 6Again, we are ignoring corrections due to the fact that i is necessarily different from j . (3.10) r = (3.11) 3.5 Correlations 33 More generally, one can understand the emergence of these correlations in the following way. For the network to become heterogeneous, we must have π(k) + N −1 ≥ σ(k) + k/(hkiN ) for large enough k , so that highly connected nodes do not lose more edges than they can acquire (see Section 3.2). This implies that π(k) must be increasing and approximately linear or superlinear. The expected value of the degree of a node i, chosen according to π(ki ), is then E (ki ) = N −1 Pk π(k)k & hk2i/hki, while that of its new, randomly chosen neighbour, j , is only E (kj ) = hki. This induces disassortative correlations which can never be compensated by the breaking of edges between nodes whose expected degree values are N −1 Pk σ(k)k and hk2i/hki if σ(k) is an increasing function. It thus ensues that a scenario such as the one analysed in this paper will never lead to assortative networks except for some cases in which σ(k) is a decreasing function – meaning that less connected nodes should be more likely to lose edges. Assortativity could, however, arise if there were some bias also on the node chosen to be i’s neighbour, e.g. on the postsynaptic neuron – which is precisely what happens in most social networks, where individuals do not generally choose their friends, partners, etc. randomly. Although there seem to be other reasons for the ubiquity of disassortative networks in nature (Johnson et al., 2010b), it is possible that the generality of the scenario studied here may also play a part. We can use the expected value matrix ǫ to estimate other magnitudes. For example, the clustering coefficient, as defined by Watts and Strogatz (Watts and Strogatz, 1998), is an average over nodes of Ci , with Ci the pro- portion of i’s neighbours which are connected to each other; so its expected value is E (Ci ) = ǫj l conditioned to j and l being neighbours of i’s. This means that, on average, we can make the approximation that kj = kl = hknn i = hki [hkihkα−1i + hkα+1 ihk−1i]. 2hkαi Substituting this value in Eq. (3.10), and taking into account that one edge of j ’s and one of l’s are taken up by i, we have C ≃ hki (hknn i − 1)α . hkαiN For a rough estimate of the mean minimum path (the minimum path between two nodes being the smallest number of edges one has to follow to get from one to the other), we can proceed as Albert and Barab´asi (2002). For a given node, let us define the number of nearest neighbours, z1 , next-nearest neighbours, z2 , and in general mth neighbours, zm . Using the relation zm = z1 (z2/z1)m−1 , (3.12) Chapter 3. Evolving networks and the development of neural systems 34 and assuming that the network is connected and can be obtained in l steps, this yields 1 + zm = N . (3.13) l X1 On average, z1 = hki and z2 = hki[(1 − C )hknni − 1] (since for each second nearest neighbour, one edge goes to the reference node and a proportion C to mutual neighbours). Now, if N ≫ z1 and z2 ≫ z1 , Eq. (3.13) leads to ln(N/hki) ln[(1 − C )hknni − 1] l ≃ 1 + (3.14) . 3.6 The C. Elegans neural network There exists a biological neural network which has been entirely mapped (although not, to the best of our knowledge, at different stages of develop- ment) – that of the much-investigated worm C. Elegans (White et al., 1986; Watts and Strogatz, 1998). With a view to testing whether such a network could arise via simple stochastic rules of the kind we are here considering, we ran simulations for the same number of nodes, N = 307, and (stationary) mean degree, hki = 14.0 (in the simple, undirected representation of the network). Using the global probabilities given by Eq. (3.6) and local ones π(k) ∼ kα and σ(k) ∼ k (as in Fig. 3.3), we obtain a surprising result. Precisely at the critical point, α = αc ≃ 1.35, there are some remarkable similarities between the biological network and the ones produced by the model. Figure 3.4 displays the degree distributions, both for the empirical net- work and for the average (stationary) simulated network corresponding to the critical point, while the top inset shows the mean-nearest-neighbour degree function knn(k) for the same networks. Both p(k) and knn(k) of the simulated networks can be seen to be very similar to those measured in the biological one. Furthermore, as is displayed in Table 3.1, the clustering coefficient obtained in simulation is almost the same as the empirical one. The mean minimum path is similar though slightly smaller in simulation, probably due to the worm’s brain having modules related to functions (Arenas et al., 2008b). Finally, Pearson’s coefficient is also in fairly good agreement, although the simulated networks are actually a bit more disassortative. It should, however, be stressed that the simulation results are for averages over 100 runs, while the biological system is equivalent to a single run; given the small number of neurons, statistical fluc- tuations can be fairly large, so one should refrain from attributing too much 3.6 The C. Elegans neural network 35 0.1 0.1 0.1 0.01 0.01 0.01 0.001 0.001 0.001 ) ) ) k k k ( ( ( p p p 1 1 m m 0 0 1e-04 1e-04 1e-04 1 1 1 102 102 102 102 n n n n n n n n k k k k 101 101 101 101 100 100 100 100 k-0.5 k-0.5 k-0.5 k-0.5 102 102 102 102 k k k k k-2.5 k-2.5 k-2.5 α α c=1.35 c=1.35 0.5 1 1.5 2 0.5 1 1.5 2 α α 10 10 10 100 100 100 1000 1000 1000 k k k Figure 3.4: Degree distribution (binned) of the C. Elegans neural network (circles) (White et al., 1986) and that obtained with MC simulations (line) in the stationary state (t = 105 steps) for an equivalent network in which edges are removed randomly (β = 1) at the critical point (α = 1.35). N = 307, κst = 14.0, averages over 100 runs. Global probabilities as in Eq. (3.6). The slope is for k−5/2 . Top right inset: mean-neighbour-degree function knn (k) as measured in the same empirical network (circles) and as given by the same simulations (line) as in the main panel. The slope is for k−1/2 . Bottom left inset: mst of equivalent network for a range of α, both from simulations (circles) and as obtained with Eq. (3.1). (See also Table 3.1.) importance to the precise values obtained – at least until we can average over 100 worms. Table 3.1 also shows the values of C , l and r both as estimated form the theory laid out in Section 3.5, and for the equivalent network in the configuration model (Newman, 2003c) – generally taken as the null model for heterogeneous networks, where the probability of an edge existing between nodes i and j is kikj /(hkiN ). It is clear that whereas the configuration-model predictions deviate substantially from the magnitudes measured in the C. El- egans neural network, the growth process we are here considering accounts for them quite well. It is interesting that it should be at the critical point that a structural topology so similar to the empirical one emerges, since it seems that the brain’s functional topology may also be related to a critical point (Chialvo, 2004; Chialvo et al., 2008). Chapter 3. Evolving networks and the development of neural systems 36 Experiment Simulation Theory Config. C 0.28 2.46 l r 0.28 2.19 0.23 0.15 1.86 1.96 -0.163 -0.207 -0.305 -0.101 Table 3.1: Values of small-world parameters C and l, and Pearson’s correla- tion coefficient r , as measured in the neural network of the worm C. Elegans (White et al., 1986), and as obtained from simulations in the stationary state (t = 105 steps) for an equivalent network at the critical point when edges are removed randomly – i.e., for α = 1.35 and β = 1. N = 307, κst = 14.0; averages over 100 runs and global probabilities as in Eq. (3.6). Theoretical estimates correspond to Eqs. (3.12), (3.14) and (3.11) applied to the networks generated by the same simulations. The last column lists the respective con- figuration model values: C and l are obtained theoretically as in (Newman, 2003c), while r , from MC simulations as in (Maslov et al., 2004), is the value expected due to the absence of multiple edges. (See also Fig. 3.4.) 3.7 Discussion With this work we have attempted, on the one hand, to extend our under- standing of evolving networks so that any choice of transition probabilities dependent on local and/or global degrees can be treated analytically, thereby obtaining some model-independent results; and on the other, to illustrate how such a framework can be applied to realistic biological scenarios. For the latter, we have used two examples relating to two rather different nervous systems: i) synaptic pruning in humans, for which the use of nonlinear global prob- abilities reproduces the initial increase and subsequent depletion in synaptic density in good accord with experiments – to the extent that nonmonotonic data points spanning a lifetime can be very well fitted with only two parame- ters; and ii) the structure of the C. Elegans neural network, for which it turns out that by only considering the numbers of nodes and edges, and imposing random deletion of edges and power-law probability of growth, the critical point leads 3.7 Discussion 37 to networks exhibiting many of the worm’s nontrivial features – such as the degree distribution, small-world parameters, and even level of disassortativity. These examples indicate that it is not far-fetched to contemplate how many structural features of the brain or other networks – and not just the degree distributions – could arise by simple stochastic rules like the ones con- sidered; although, undoubtedly, other ingredients such as natural modularity (Arenas et al., 2008b), a metric (Kaiser and Hilgetag, 2004) or functional re- quirements (Sporns et al., 2004) can also be expected to play a role in many instances. We hope, therefore, that the framework laid out here – in which for simplicity we have assumed the network to be undirected and to have a fixed size, although generalizations are straightforward – may prove useful for interpreting data from a variety of fields. It would be particularly interesting to try to locate and quantify the biological mechanisms assumed to be behind this kind of network dynamics. Chapter 4 Bringing on the Edge of Chaos with heterogeneity The collective behaviour of systems of coupled excitable elements, such as neurons, has been shown to depend significantly on the heterogeneity of the degree distribution of the underlying network of interactions. For instance, broad – in particular, scale-free – distributions have been found to improve static memory performance in neural-network models. Here we look at the influence of degree heterogeneity in a neural network which, due to the effect of synaptic depression (a kind of fatigue of the interaction strengths), exhibits chaotic behaviour. Not only can the existence of a chaotic phase be related to neurophysiological experiments; it allows the system to perform a class of dynamic pattern-recognition tasks. We find first of all that, as has been described in a few other systems, optimal performance is achieved close to the phase transition – i.e., at the so-called Edge of Chaos. Furthermore, we obtain a functional relationship between the level of synaptic depression required to bring on chaos and the heterogeneity of the degree distribution. This result points to a clear advantage of low-exponent scale-free networks, and suggests an explanation for their apparent ubiquity in certain biological systems. 4.1 Exciting cooperation Excitable systems allow for the regeneration of waves propagating through them, and may thus respond vigorously to weak stimulus. The brain and other parts of the nervous system are well–studied paradigms, and forest fires with constant ignition of trees and autocatalytic reactions in surfaces, for instance, also share some of the basics (Bak et al., 1990; Meron, 1992; Lindner et al., 39 40 Chapter 4. Bringing on the Edge of Chaos with heterogeneity 2004; Izhikevich , 2007; Arenas et al., 2008a). The fact that signals are not gradually damped by friction in these cases is a consequence of cooperation among many elements in a nonequilibrium setting. These systems can be seen as large networks of nodes that are “excitable”. This admits various realizations, but typically means that each element has a threshold and a refractory time between consecutive responses – a behaviour that impedes thermal equilibrium. Some brain tasks can be simulated with mathematical neural networks. As described in Chapter 2, these consist of neurons – often modelled as variables which are as simple as possible while still able to display the essence of the cooperative behaviour of interest1 – connected by edges representing synapses (Amari, 1972; Hopfield, 1982; Amit, 1989; Torres and Varona, 2010). If the edges are weighted according to some prescription – such as the Hebb rule (Hebb, 1949) – which saves information from a set of given patterns of activ- ity (particular configurations of active and inactive neurons), these patterns become attractors of the phase-space dynamics. Therefore, the system is then able to retrieve the stored patterns; this mechanism is known as associative memory. Actual neural systems do much more than just recalling a memory and staying there, however. That is, one should expect dynamic instabilities or some other destabilizing mechanism. This expectation is reinforced by re- cent experiments suggesting that synapses undergo rapid changes with time which may both determine brain tasks (Abbott et al., 1997; Tsodyks et al., 1998; Hilfiker et al., 1999; Pantic et al., 2002) and induce irregular and per- haps chaotic activity (Barrie et al., 1996; Korn and Faure, 2003). One may argue that the observed rapid changes (which have been found to cause “synaptic depression” and/or “facilitation” on the time scale of mil- liseconds (Tsodyks et al., 1998; Pantic et al., 2002) – i.e., much faster than the plasticity processes whereby synapses store patterns (Malenka and Nicoll, 1999)) may simply correspond to the characteristic behaviour of single ex- citable elements. Furthermore, a fully-connected network which describes co- operation among such excitable elements has recently been shown to exhibit both attractors and chaotic instabilities (Marro et al., 2008). The work de- scribed here, first reported by Johnson et al. (2008), extends and generalizes this study to conclude on the influence of the excitable network topology on 1Several studies have already shown that binary neurons can capture the essence of cooperation in many more complex settings. See, for instance, (Pantic et al., 2002) in the case of integrate and fire neuron models of pyramidal cells. 4.2 The Fast-Noise model 41 dynamic behaviour. We show, in particular, an interesting correlation between certain wiring topology and optimal functionality. 4.2 The Fast-Noise model Consider N binary nodes (si = ±1) and the adjacency matrix, aij = 1, 0, which indicates the existence or not of an edge between nodes i, j = 1, 2, ..., N . Let there be a set of M patterns, ξ ν i = ±1, ν = 1, ...M (which we generate here at random), and assume that they are “stored” by giving each edge a base weight ωij = N −1 Pν ξ ν i ξ ν j . Actual weights are dynamic, however, namely, ωij = ωij xj where xj is a stochastic variable. Assuming the limit in which this varies in a time scale infinitely smaller than the one for node dynamics, we can consider a stationary distribution such as P (xj S ) = qδ (xj − Ξj ) + (1 − q)δ (xj − 1), S = {sj } , for instance. This amounts to assuming that, at each time step, every connection has a probability q of altering its weight by a factor Ξj which is a function (to be determined) of the local field at j, defined as the net current arriving to j from other nodes. This choice differs essentially from the one used by Marro et al. (2008), where q depends on the global degree of order and Ξj is a constant independent of j. Assume independence of the noise at different edges, and that the transition rate for the stochastic changes is = Yj /aij =1 R dxj P (xj S )Ψ(uij ) ¯c (S → S i ) R dxj P (xj S i)Ψ(−uij ) ¯c (S i → S ) where uij ≡ sisj xj ωij T −1 , Ψ(u) = exp (cid:0)− 1 2 u(cid:1) to have proper contour condi- tions, T is a “temperature” or stochasticity parameter, and S i stands for S after the change si → −si . (This formalism and its interpretation is described in detail by Marro and Dickman.) We define the effective local fields heff i = i si /T (cid:1) , where ϕ± i (S, T , q) via Qj ϕ− ij /ϕ+ heff ij = exp (cid:0)−heff ij ≡ q exp (±Ξj vij ) + (1 − q) exp (±vij ), with vij = 1 2 aij uij . Effective weights ω eff ij then follow from heff i = Pj ω eff ij sj aij . To obtain an analytical expression, we linearize around ωij = 0 (a good approximation when M ≪ N ), which yields ω eff ij = [1 + q (Ξj − 1)] ωij . In order to fix Ξj here, we first introduce the overlap vector −→m = (m1 , ...mM ), with mν ≡ N −1 Pi ξ ν i si , which measures the correlation between the current configuration and each of the stored patterns, and the local one −→mj of com- ponents mν j ≡ hki−1 Pl ξ ν l sl aj l , where hki is the mean node connectivity, i.e., , 42 Chapter 4. Bringing on the Edge of Chaos with heterogeneity the average of ki = Pj aij . We then assume, for any q 6= 0, that the relevant factor is Ξj = 1 + ζ (hν j )(Φ − 1)/q , with χα 1 + M/N Xν ζ (hν hν j α , j ) = where χ ≡ N/hki and α > 0 is a parameter. This comes from the fact that the field at node j can be written as a sum of components from each pattern, namely, hj = PM ν hν j , where j N −1 Xi j = ξ ν hν Our choice for Ξj , which amounts to assuming that the “fatigue” at a given edge increases with the field at the preceding node j (and allows to recover the fully–connected limit described by Marro et al. (2008) if α = 2), finally leads to j mν i si = χ−1ξ ν aij ξ ν j . ij = [1 + (Φ − 1)ζj (−→mj )] ωij . ω eff Varying Φ one sets the nature of the weights. That is, 0 < Φ < 1 corresponds to resistance (depression ) due to heavy local work, while the edge facilitates – i.e., tends to increase the effect of the signal under the same situation – for Φ > 1. (The action of the edge is reversed for negative Φ.) We performed Monte Carlo simulations using standard parallel updating with the effective rates ¯c (S → S i ) computed using the latter effective weights. 4.3 Edge of Chaos It is possible to solve the single pattern case (M = 1) under a mean-field as- sumption, which is a good approximation for large enough connectivity. That is, we may substitute the matrix aij by its mean value over network realizations to obtain analytical results that are independent of the underlying disorder. Imagine that each node hosts ki half–edges according to a distribution p(k), the total number of half–edges in the network being hkiN . Choose a node i at random and randomly join one of its half–edges to an available free half– edge. The probability that this half–edge ends at node j is kj / (hki N ) . Once all the nodes have been linked up, the expected value (as a quenched average over network realizations) for the number of edges joining nodes i and j is2 2Assuming one edge at most between any two nodes, aij = 0, 1, the value will be slightly smaller, but it is easy to prove that this is also a good approximation if the network has a structural cut-off: ki < phkiN , ∀i. 4.3 Edge of Chaos 43 E (aij ) = kikj / (hki N ). This expression, which can be seen as a definition of the so-called configuration model for complex networks (Newman, 2003c), is valid for random networks with a given degree sequence (or, in practise, a given degree distribution) that have zero degree-degree correlations between neighbours (Johnson et al., 2010b). Using the notation ηi ≡ ξisi , we have mj = χhηi aij ii = χ N Pi ηi aij . Because node activity is not statistically in- dependent of connectivity (Torres et al., 2004), we must define a new set of overlap parameters, analogous to m and mj . That is, µn ≡ hkn i ηi ii/hkni and the local versions µj n ≡ χhkn i ηi aij ii/hkni. After using aij = E (aij ), one ob- tains the relation µi n = hkn+1ikiµn+1/(hknihki2). Inserting this expression into the definition of µn , and substituting hsi i = tanh[T −1hef f (S )] (for large N), i standard mean-field analysis yields MT ,Φ = µn (t + 1) = 1 hkn i hkn tanh MT ,Φ (k , t)ik , where the last quantity is defined as T N (cid:20)µ1(t) + (Φ − 1) hkα+1 i hkiα+1 µ1(t)α µα+1(t)(cid:21) . k This is a two-dimensional map which is valid for any random topology of distribution p(k). Note that the macroscopic magnitude of interest is µ0 = m ≡ −→m . A main consequence of this is the existence of a critical temperature, Tc , un- der very general conditions. More specifically, as T is decreased, the overlap m describes a second–order phase transition from a disordered or, say, “paramag- netic” phase to an ordered (“ferromagnetic”) phase which exhibits associative memory. The mean-field temperature at which this transition occurs is Tc = hk2i hki N On the other hand, the map reduces to µn (t + 1) = sign (cid:26)µn (t) (cid:20)1 + (Φ − 1) hkα+1i hkiα+1 (cid:21)(cid:27) for T = 0. This implies the existence at Φ = Φ0 , where Φ0 = 1 − hkiα+1 hkα+1i of a transition as Φ is decreased from the ferromagnetic phase to a new phase in which periodic hopping between the attractor and its negative occurs. This . , 44 Chapter 4. Bringing on the Edge of Chaos with heterogeneity is confirmed by the Monte Carlo simulations for M > 1; that is, the hopping is also among different attractors for finite T . The simulations also indicate that this transition washes out at low enough finite temperature. Instead, Monte Carlo evolutions show that, for a certain range of Φ values, the system activity then exhibits chaotic behaviour. The transition from ferromagnetic to chaotic states is a main concern here- after. Our interest in this regime follows from several recent observations concerning the relevance of chaotic activity in a network. In particular, it has been shown that chaos might be responsible for certain states of attention dur- ing brain activity (Torres et al., 2008, 2009), and that some network properties such as the computational capacity (Bertschinger and Natschlager, 2004) and the dynamic range of sensitivity to stimuli (de Assis and Copelli, 2008) may become optimal at the Edge of Chaos in a variety of settings. We next note that the critical values Tc and Φ0 only depend on the mo- ments of the generic distribution p(k), and that the ratio hkai/hkia , a > 1, is a convenient way of characterizing heterogeneity. We studied in detail two particular types of connectivity distributions with easily tunable hetero- geneity; that is, networks with hkiN/2 edges randomly distributed with p (k) such that the heterogeneity depends on a single parameter. Our first case 2 δ (k − k1) + 1 is the bimodal distribution, p(k) = 1 2 δ (k − k2) with parameter ∆ = (k2 − k1)/2 = hki − k1 = k2 − hki. Our second case is the scale–free distribution, p(k) ∼ k−γ , which does not have any characteristic size but k is 1 γ−1 , N − 1) for finite N . Notice confined to the limits, k0 and km ≤ min(k0N that the network in this case gets more homogeneous as γ is increased3 , and that this kind of distribution seems to be most relevant in nature (Newman, 2003c; Boccaletti et al., 2006). In particular, it seems important to mention that the functional topology of the human brain, as defined by correlated ac- tivity between small clusters of neurons, has been shown to correspond to this case with exponent γ ≃ 2 (Egu´ıluz et al., 2005). (It has not yet been possible to ascertain the brain’s structural topology experimentally, but there is some evidence that function reflects structure at least to some extent (Zhou et al., 2006b). Furthermore, it has been suggested, based on indirect methods, that the structural connectivity of cat and macaque brains, at the level of brain areas, may indeed be scale free (Kaiser et al., 2007) – and in any case dis- 3The distribution is truncated and therefore not strictly scale free for γ < 2. However, nature shows examples for which γ is slightly larger than 1, so we consider the whole range here. 4.3 Edge of Chaos 45 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 ∆=4 ∆=4 ∆=4 ∆=4 ∆=8 ∆=8 ∆=8 ∆=12 ∆=12 γ=1.5 γ=1.5 γ=1.5 γ=1.5 γ=2.5 γ=2.5 γ=2.5 γ=3.5 γ=3.5 c c c c c c c c Φ Φ Φ Φ Φ Φ Φ Φ - - - - - - - - 0 0 0 0 0 0 0 0 Φ Φ Φ Φ Φ Φ Φ Φ 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 T/Tc T/Tc T/Tc T/Tc T/Tc T/Tc T/Tc T/Tc Figure 4.1: The temperature dependence of the difference between the values for the fatigue at which the ferromagnetic–periodic transition occurs, as ob- tained analytically for T = 0 (Φ0 ) and from MC simulations at finite T (Φc ). The critical temperature is calculated as Tc = hk2 i (hki N )−1 for each topology. Data are for bimodal distributions with varying ∆ and for scale–free topologies with varying γ , as indicated. Here, hki = 20, N = 1600 and α = 2. Standard deviations, represented as bars in this graph, were shown to drop with N −1/2 (not depicted). plays significantly higher heterogeneity than that of, say, Erdos–R´enyi random graphs.) We obtained the critical value of the fatigue, Φc (T ) , from Monte Carlo simulations at finite temperature T . These indicate that chaos never occurs for T & 0.35Tc . On the other hand, a detailed comparison of the value Φc with Φ0 – as obtained analytically for T = 0 – indicates that Φc ≃ Φ0 . Figure 4.1 illustrates the “error” Φ0 − Φc (T ) for different topologies. This shows that the approximation Φc ≃ Φ0 is quite good at low T for any of the cases examined. Therefore, assuming the critical values for the main param- eters, Tc and Φ0 , as given by our map, we conclude that the more heteroge- neous the distribution of connectivities of a network is, the lower the amount of fatigue, and the higher the critical temperature, needed to destabilize the dynamics. As an example of this interesting behaviour, consider a network with hki = ln(N ), and dynamics according to α = 2. If the distribution were 46 Chapter 4. Bringing on the Edge of Chaos with heterogeneity 0.5 0.5 0.5 Stable Stable Stable Φ Φ Φ Unstable Unstable Unstable 0 0 0 1 1 1 1 1 1 0 0 0 4 4 4 8 8 8 ∆ ∆ ∆ 12 12 12 16 16 16 Stable Stable Stable Stable Stable Stable Φ Φ Φ Φ Φ Φ 0.5 0.5 0.5 0.5 0.5 0.5 Unstable Unstable Unstable Unstable Unstable Unstable 0 0 0 0 0 0 1 1 1 1 1 1 2 2 2 2 2 2 3 3 3 3 3 3 4 4 4 4 4 4 5 5 5 5 5 5 6 6 6 6 6 6 γ γ γ γ γ γ Figure 4.2: The critical fatigue values Φ0 (solid lines) and Φc from MC averages over 10 networks (symbols) with T = 2/N , hki = 20, N = 1600, α = 2. The dots below the lines correspond to changes of sign of the Lyapunov exponent as given by the iterated map, which qualitatively agree with the other results. This is for bimodal and scale–free topologies, as indicated. regular, the critical values would be Tc = ln(N )/N (which goes to zero in the thermodynamic limit) and Φ0 = 0. However, a scale–free topology with the same number of edges and γ = 2 would yield Tc = 1 and Φ0 = 1 − 2(ln N )3/N 2 (which goes to 1 as N → ∞). Figure 6.5 illustrates, for two topologies, the phase diagram of the ferromagnetic– chaotic transition. Most remarkable is the plateau observed in the Edge-of- 4.4 Network performance 47 Chaos or transition curve for scale–free topologies around γ ≃ 2, for which very little fatigue, namely, Φ . 1 which corresponds to slight depression, is required to achieve chaos. The limit γ → ∞ corresponds to hki–regular graphs (equivalent to ∆ = 0). If γ is reduced, km increases and k0 decreases. The network is truncated when km = N . It follows that a value of γ exits at which k0 cannot be smaller, so that km must drop to preserve hki. This explains the fall in Φc as γ → 1. Assuming that the “ferromagnetic phase” here corresponds to a synchronous state, our results are in qualitative agreement with the ones obtained recently for coupled oscillators (Nishikawa et al., 2003; Zhou et al., 2006a). As a matter of fact, the range of coupling strengths which allow for stability of synchronous states in these systems has been shown to depend on the spectral gap of the Laplacian matrix (Barahona and Pecora, 2002), implying that the more het- erogeneous a topology is, the more easily activity can become unstable. It should be emphasized, however, that the dynamics we are considering here does not come within the scope of the formalism used to derive these results, since activity at node i depends on the local field at node j . 4.4 Network performance As a further illustration of our findings, we monitored the performance as a function of topology during a simulation of pattern recognition. That is, we “showed” the system a pattern, say ν chosen at random from the set of M previously stored, every certain number of time steps. This was performed in practice by changing the field at each node for one time step, namely, hi → hi + δξ ν , where δ measures the intensity of the input signal. Ideally, the network should remain in this configuration until it is newly stimulated. The performance may thus be estimated from a temporal average of the overlap between the current state and the input pattern, hmν itime . This is observed to simply increase monotonically with ∆ for the bimodal case. The scale–free case, however, as illustrated in Fig. 4.3, shows how the task is better performed the closer to the Edge of Chaos the network is. This is because the system is then easily destabilized by the stimulus while being able to retrieve a pattern with accuracy. Figure 4.3 also shows that the best performance for the scale– free topology when Φ = 1, i.e., in the absence of any fatigue, definitely occurs around γ = 2. 48 Chapter 4. Bringing on the Edge of Chaos with heterogeneity 1 1 Φ=0.8 Φ=0.8 t t > > ν ν m m < < 0.5 0.5 0 0 1 1 5 5 Φ=1.0 Φ=1.0 t t > > ν ν m m < < 0.5 0.5 1 1 2 2 1 1 1 1 0 0 0 0 1 1 1 1 0 0 0 0 ∆ ∆ 1 1 1 1 0 0 0 0 1 1 1 1 0 0 0 0 3 3 γ γ Time Time Time Time 10 10 15 15 Time Time Time Time 4 4 5 5 Figure 4.3: Network “performance” (see the main text) against ∆ for bimodal topologies (above) and against γ for scale–free topologies (below). Φ = 0.8 for the first case and Φ = 1 in the second. Averages over 20 network realizations with stimulation every 50 MC steps for 2000 MC steps, δ = 5 and M = 4; other parameters as in Fig. 6.5. Inset shows sections of typical time series of mν for ∆ = 10 (above) and γ = 4 (below); the corresponding stimulus for pattern ν is shown underneath. 4.5 Discussion The model network we have studied is one of the simplest relevant situations one may conceive. In particular, as emphasized above, we are greatly simply- fieng the elements at the nodes (neurons) as binary variables. However, our 4.5 Discussion 49 assumption of dynamic connections which depend on the local fields in such a simple scenario happens to show that a close relation may exist between topological heterogeneity and function, thus suggesting this may indeed be a relevant property for a realistic network efficiently to perform certain high level tasks. In a similar way to networks shown previously to be useful for pattern recognition and family identification (Cortes et al., 2005), our system retrieves memory patterns with accuracy in spite of noise, and yet it is easily destabi- lized so as to change state in response to an input signal – without requiring excessive fatigue for the purpose. There is a relation between the amount Φ of fatigue and the value of γ for which performance is maximized. One may argue that the plateau of “good” behaviour shown around γ ≃ 2 for scale–free networks with Φ . 1 (Fig. 6.5) is a possible justification for the supposed tendency of certain systems in nature to evolve towards this topology. It may also prove useful for implementing some artificial networks. Chapter 5 Correlated networks and natural disassortativity An intriguing feature of complex networks is the ubiquity of strong negative degree-degree correlations between neighbouring nodes – the only exceptions being social systems, which tend to be assortative instead of disassortative. With the double purpose of addressing this mystery and uncovering the effects of correlations on network behaviour, we put forward a method which allows for the model-independent study of ensembles of correlated networks. We go on to show, by means of an information theory approach, that the expected value of correlations for a network at equilibrium (i.e., in the absence of spe- cific correlating mechanisms) is not, as had been supposed, uncorrelated, but rahter disassortative. It turns out that the correlations of some networks are in excellent agreement with our predictions, while others, with known correlating or anticorrelating mechanisms, indeed appear to have been driven from their equilibrium points as expected. Therefore, our approach not only provides a parsimonious topological answer to a long-standing question, but also a neu- tral model against which to contrast experimental data to determine whether mechanisms must be sought to account for observed correlations. We go on to use our method, in Chapter 6, to study the influence of assortativity on neural-network dynamics. 5.1 Assortativity of networks Complex networks, whether natural or artificial, have non-trivial topologies which are usually studied by analysing a variety of measures, such as the degree distribution, clustering, average paths, modularity, etc. (Albert and Barab´asi, 51 52 Chapter 5. Correlated networks and natural disassortativity 2002; Dorogovtsev and Mendes, 2003; Pastor-Satorras and Vespignani, 2004; Newman, 2003c; Boccaletti et al., 2006) The mechanisms which lead to a par- ticular structure and their relation to functional constraints are often not clear and constitute the sub ject of much debate (Newman, 2003c; Boccaletti et al., 2006). When nodes are endowed with some additional “property,” a feature known as mixing or assortativity can arise, whereby edges are not placed be- tween nodes completely at random, but depending in some way on the property in question. If similar (dissimilar) nodes tend to wire together, the network is said to be assortative (disassortative ) (Newman, 2002, 2003a). An interesting situation is when the property taken into account is the degree of each node – i.e., the number of neighbouring nodes connected to it. It turns out that a high proportion of empirical networks – whether biological, technological, information-related or linguistic – are disassortatively arranged (high-degree nodes, or hubs, are preferentially linked to low-degree neighbours, and viceversa) while social networks are usually assortative. Such degree- degree correlations have important consequences for network characteristics such as connectedness and robustness (Newman, 2002, 2003a). However, while assortativity in social networks can be explained taking into account homophily (Newman, 2002, 2003a) or modularity (Newman and Park, 2003), the widespread prevalence and extent of disassortative mixing in most other networks remains somewhat mysterious. Maslov et al. found that the restriction of having at most one edge per pair of nodes induces some dis- assortative correlations in heterogeneous networks (Maslov et al., 2004), and Park and Newman showed how this analogue of the Pauli exclusion principle leads to the edges following Fermi statistics (Park and Newman, 2003) (see also (Capocci and Colaiori, 2006)). However, this restriction is not sufficient to fully account for empirical data. In general, when one attempts to consider computationally all the networks with the same distribution as a given empiri- cal one, the mean assortativity is not necessarily zero (Holme and Zhao, 2007). But since some “randomization” mechanisms induce positive correlations and others negative ones (Farkas et al., 2004; Johnson et al., 2010a), it is not clear how the phase space can be properly sampled numerically. In this chapter we develop a method for the study of correlated networks which is model-independent, and describe the main result of Ref. (Johnson et al., 2010b) – namely, that there is a general reason, consistent with empirical data, for the “natural” mixing of most networks to be disassortative. Us- ing an information-theory approach we find that the configuration which can 5.2 The entropy of network ensembles 53 be expected to come about in the absence of specific additional constraints turns out not to be, in general, uncorrelated. In fact, for highly heterogeneous degree distributions such as those of the ubiquitous scale-free networks, we show that the expected value of the mixing is usually disassortative: there are simply more possible disassortative configurations than assortative ones. This result provides a simple topological answer to a long-standing question. Let us caution that this does not imply that all scale-free networks are disassortative, but only that, in the absence of further information on the mechanisms behind their evolution, this is the neutral expectation. 5.2 The entropy of network ensembles The topology of a network is entirely described by its adjacency matrix a; the element aij represents the number of edges linking node i to node j (for undi- rected networks, a is symmetric). Among all the possible microscopically dis- tinguishable configurations a set of L edges can adopt when distributed among N nodes, it is often convenient to consider the set of configurations which have certain features in common – typically some macroscopic magnitude, like the degree distribution. Such a set of configurations defines an ensemble. In a seminal series of papers Bianconi has determined the partition functions of various ensembles of random networks and derived their statistical-mechanics entropy (Bianconi, 2008, 2009; Anand and Bianconi, 2009). This allows the author to estimate the probability that a random network with certain con- straints has of belonging to a particular ensemble, and thus assess the relative importance of different magnitudes and help discern the mechanisms respon- sible for a given real-world network. For instance, she shows that scale-free networks arise naturally when the total entropy is restricted to a small finite value. Here we take a similar approach: we obtain the Shannon information entropy encoded in the distribution of edges. As we shall see, both methods yield the same results (Jaynes, 1957; Anand and Bianconi, 2009), but for our purposes the Shannon entropy is more tractable. The Shannon entropy associated with a probability distribution pm is s = − Xm where the sum extends over all possible outcomes m. For a given pair of nodes (i, j ), pm can be considered to represent the probability of there being m edges between i and j . For simplicity, we shall focus here on networks such pm ln(pm), 54 Chapter 5. Correlated networks and natural disassortativity that aij can only take values 0 or 1, although the method is applicable to any number of edges allowed. In this case, we have only two terms: p1 = ǫij and p0 = 1 − ǫij , where ǫij ≡ E (aij ) is the expected value of the element aij given that the network belongs to the ensemble of interest. The entropy associated with pair (i, j ) is then (5.1) sij = − [ǫij ln(ǫij ) + (1 − ǫij ) ln(1 − ǫij )] , while the total entropy of the network is S = PN ij sij : N Xij S = − [ǫij ln(ǫij ) + (1 − ǫij ) ln(1 − ǫij )] . Since we have not imposed symmetry of the adjacency matrix, this expression is in general valid for directed networks. For undirected networks, however, the sum is only over i ≤ j , with the consequent reduction in entropy. For the sake of illustration, we shall estimate the entropy of the Internet at the autonomous system (AS) level and compare it with the values obtained in (Bianconi, 2008, 2009; Anand and Bianconi, 2009) assuming the network belongs to two different ensembles: the fully random graph, or Erdos-R´enyi (ER) ensemble, and the configuration ensemble with a scale-free degree distri- bution (p(k) ∼ k−γ ) (Newman, 2003c) and structural cutoff, ki < phkiN , ∀i (Bianconi, 2008, 2009; Anand and Bianconi, 2009) (hki is the mean degree). In this example, we assume the network to be sparse enough to expand the term ln(1 − ǫij ) in Eq. (5.1) and keep only linear terms. This reduces Eq. (5.1) to N Xij In the ER ensemble, each of N nodes has an equal probability of receiving each of 1 2 hkiN undirected edges. So, writing ǫER ij = hki/N , we have 1 2 hkiN [ln (hki/N ) − 1] . SER = − The configuration ensemble, which imposes a given degree sequence (k1 , ...kN ), is defined via the expected value of the adjacency matrix (Newman, 2003c; Johnson et al., 2008): ǫij [ln(ǫij ) − 1] + O(ǫ2 ij ). Ssparse ≃ − This value leads to ǫc ij = kikj /(hkiN ). Sc = hkiN [ln(hkiN ) + 1] − 2N hk ln ki, 5.3 Entropic origin of disassortativity 55 SB ER SB c SER Sc(γ=2.3) 20 20 20 20 20 20 20 20 15 15 15 15 15 15 15 15 10 10 10 10 10 10 10 10 e e e e e e e e d d d d d d d d o o o o o o o o n n n n n n n n r r r r r r r r e e e e e e e e p p p p p p p p y y y y y y y y p p p p p p p p o o o o o o o o r r r r r r r r t t t t t t t t n n n n n n n n E E E E E E E E 11/97 11/97 11/97 11/97 11/97 11/97 11/97 11/97 10/98 10/99 10/99 10/99 10/99 10/99 10/99 10/99 10/99 10/98 10/98 10/98 10/98 10/98 10/98 10/98 Date (month/year) Date (month/year) Date (month/year) Date (month/year) Date (month/year) Date (month/year) Date (month/year) Date (month/year) 10/00 10/00 10/00 10/00 10/00 10/00 10/00 10/00 3/01 3/01 3/01 3/01 3/01 3/01 3/01 3/01 Figure 5.1: Evolution of the Internet at the AS level. Empty (blue) squares and circles: entropy per node of randomized networks in the fully random and in the configuration ensembles, as obtained by Bianconi (hence the “B” su- perscription) (Bianconi, 2008, 2009; Anand and Bianconi, 2009). Filled (red) triangles and diamonds: Shannon entropy for an ER network and a scale-free one with γ = 2.3, respectively. where h·i ≡ N −1 Pi (·) stands for an average over nodes. Fig. 5.1 displays the entropy per node obtained in (Bianconi, 2008, 2009; Anand and Bianconi, 2009) for the first two levels of approximation (ensem- bles) to the Internet at the AS level, first taking into account only the numbers of nodes N and edges L = 1 2 hkiN , and then also the degree sequence. Along- side these, we plot the Shannon entropy both for an ER random network, (which coincides exactly with Bianconi’s expression), and for a scale-free net- work with γ = 2.3 (the slight disparity arising from this exponent’s changing a little with time). 5.3 Entropic origin of disassortativity We shall now go on to analyse the effect of degree-degree correlations on the entropy. In the configuration ensemble, the expected value of the mean degree 56 Chapter 5. Correlated networks and natural disassortativity Figure 5.2: Shannon entropy of correlated scale-free networks against param- eter β (left panel) and against Pearson’s coefficient r (right panel), for various values of γ (increasing from bottom to top). hki = 10, N = 104 . of the neighbours of a given node is , ij kj = hk2i i Xj ǫc knn,i = k−1 hki which is independent of ki . However, as mentioned above, real networks often display degree-degree correlations, with the result that knn,i = knn(ki ). If knn(k) increases (decreases) with k , the network is assortative (disassortative). A measure of this phenomenon is Pearson’s coefficient applied to the edges (Newman, 2003c, 2002, 2003a; Boccaletti et al., 2006): r = l ] − [kl ]2 [klk ′ l ] − [kl ]2 , [k2 where kl and k ′ l are the degrees of each of the two nodes belonging to edge l, and [·] ≡ (hkiN )−1 Pl (·) is an average over edges. Writing Pl (·) = Pij aij (·), r can be expressed as r = hkihk2knn(k)i − hk2i2 (5.2) . hkihk3i − hk2 i2 The ensemble of all networks with a given degree sequence (k1 , ...kN ) contains a subset for all members of which knn (k) is constant (the configuration ensemble), but also subsets displaying other functions knn (k). We can identify each one 5.3 Entropic origin of disassortativity 57 kj ǫij = kiknn (ki ), ∀i, and ǫij = ki , ∀i (for consistency). of these subsets (regions of phase space) with an expected adjacency matrix ǫ which simultaneously satisfies the following conditions: i) Xj ii) Xj An ansatz which fulfils these requirements is any matrix of the form N (cid:20) (kikj )ν j + hkν i(cid:21) , kikj f (ν ) + Z dν hkν i − kν i − kν hkiN where ν ∈ R and the function f (ν ) is in general arbitrary, although depending on the degree sequence it shall here be restricted to values which maintain ǫij ∈ [0, 1], ∀i, j . This ansatz yields knn(k) = hk2 i + Z dν f (ν )σν+1 (cid:20) kν−1 k (cid:21) 1 hkν i − hki (the first term being the result for the configuration ensemble), where σb+1 ≡ hk b+1i − hkihk bi. In practice, one could adjust Eq. (5.4) to fit any given func- tion knn(k) and then wire up a network with the desired correlations: it suffices to throw random numbers according to Eq. (5.3) with f (ν ) as obtained from the fit to Eq. (5.4)1 . To prove the uniqueness of a matrix ǫ obtained in this way (i.e., that it is the only one compatible with a given knn (k)) assume that there exists another valid matrix ǫ′ 6= ǫ. Writting ǫ′ ij − ǫij ≡ h(ki , kj ) = hij , then i) implies that Pj kj hij = 0, ∀i, while ii) means that Pj hij = 0, ∀i. It follows that hij = 0, ∀i, j . ǫij = (5.3) (5.4) In many empirical networks, knn (k) has the form knn (k) = A + Bkβ , with A, B > 0 (Boccaletti et al., 2006; Pastor-Satorras et al., 2001) – the mixing being assortative (disassortative) if β is positive (negative). Such a case is fitted by Eq. (5.4) if f (ν ) = C (cid:20)δ (ν − β − 1) σβ+2 − δ (ν − 1)(cid:21) , σ2 1Although, as with the configuration ensemble, it is not always possible to wire a network according to a given ǫ. 58 Chapter 5. Correlated networks and natural disassortativity r = (5.6) (5.5) with C a positive constant, since this choice yields knn (k) = hk2i + C σ2 (cid:20) kβ hki (cid:21) . 1 hkβ+1i − hki After plugging Eq. (5.5) into Eq. (5.2), one obtains: hkβ+1i (cid:18) hkihkβ+2i − hk2ihkβ+1i (cid:19) . C σ2 hkihk3i − hk2i2 Inserting Eq. (5.3) in Eq. (5.1), we can calculate the entropy of correlated networks as a function of β and C – or, by using Eq. (5.6), as a function of r . Particularizing for scale-free networks, then given hki, N and γ , there is always a certain combination of parameters β and C which maximizes the entropy; we shall call these β ∗ and C ∗ . For γ . 5/2 this point corresponds to C ∗ = 1. For higher γ , the entropy can be slightly higher for larger C . However, for these values of γ , the assortativity r of the point of maximum entropy obtained with C = 1 differs very little from the one corresponding to β ∗ and C ∗ (data not shown). Therefore, for the sake of clarity but with very little loss of accuracy, in the following we shall generically set C = 1 and vary only β in our search for the level of assortativity, r∗ , that maximizes the entropy given hki, N and γ . Note that C = 1 corresponds to removing the linear term, proportional to kikj , in Eq. (5.3), and leaving the leading non-linearity, (kikj )β+1 , as the dominant one. Fig. 5.2 displays the entropy curves for various scale-free networks, both as functions of β and of r : depending on the value of γ , the point of max- imum entropy can be either assortative or disassortative. This can be seen more clearly in Fig. 5.3, where r∗ is plotted against γ for scale-free networks with various mean degrees hki. The values obtained by Park and Newman (Park and Newman, 2003) as those resulting from the one-edge-per-pair re- striction are also shown for comparison: notice that whereas this effect alone cannot account for the Internet’s correlations for any γ , entropy considerations would suffice if γ ≃ 2.1. As shown in the inset, the results are robust in the large system-size limit. Since most networks observed in the real world are highly heterogeneous, with exponents in the range γ ∈ (2, 3), it is to be expected that these should display a certain disassortativity – the more so the lower γ and the higher hki. In Fig. 5.4 we test this prediction on a sample of empirical, scale-free net- works quoted in Newman’s review (Newman, 2003c) (p. 182). For each case, we found the value of r that maximizes S according to Eq. (5.1), after insert- ing Eq. (5.3) with the quoted values of hki, N and γ . In this way, we obtained 5.3 Entropic origin of disassortativity 59 <k>=1/2 k0 <k>=1/2 k0 <k>=1/2 k0 <k>=1/2 k0 <k>=1/2 k0 <k>=k0 <k>=k0 <k>=k0 <k>=k0 <k>=2 k0 <k>=2 k0 <k>=2 k0 <k>=4 k0 <k>=4 k0 r0=-0.189 r0=-0.189 r0=-0.189 r0=-0.189 r0=-0.189 * * * * * * r r r r r r 0.1 0.1 0.1 0.1 0.1 0.1 0 0 0 0 0 0 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.3 -0.3 -0.3 -0.3 -0.3 -0.3 2 2 2 2 2 2 0.1 0.1 0.1 0.1 * * * * r r r r 0 0 0 0 -0.1 -0.1 -0.1 -0.1 103 103 103 103 3 3 3 3 3 3 γ γ γ γ γ γ γ=2.5 γ=2.5 γ=2.5 γ=2.5 3´•104 3´•104 3´•104 3´•104 N N N N 4 4 4 4 4 4 Figure 5.3: Lines from top to bottom: r at which the entropy is maximized, r∗ , against γ for random scale-free networks with mean degrees hki = 1 2 , 1, 2 and 4 times k0 = 5.981, and N = N0 = 10697 nodes (k0 and N0 correspond to the values for the Internet at the AS level in 2001 (Park and Newman, 2003), which had r = r0 = −0.189). Symbols are the values obtained in (Park and Newman, 2003) as those expected solely due to the one-edge-per- pair restriction (with k0 , N0 and γ = 2.1, 2.3 and 2.5). Inset: r∗ against N for networks with fixed hki/N (same values as the main panel) and γ = 2.5; the arrow indicates N = N0 . the expected assortativity for six networks, representing: a peer-to-peer (P2P) network, metabolic reactions, the nd.edu domain, actor collaborations, protein interactions, and the Internet (see (Newman, 2003c) and references therein). For the metabolic, Web domain and protein networks, the values predicted are in excel lent agreement with the measured ones; therefore, no specific anticor- relating mechanisms need to be invoked to account for their disassortativity. In the other three cases, however, the predictions are not accurate, so there must be additional correlating mechanisms at work. Indeed, it is known that small routers tend to connect to large ones (Pastor-Satorras et al., 2001), so one would expect the Internet to be more disassortative than predicted, as is the case2 – an effect that is less pronounced but still detectable in the more 2However, as Fig. 5.3 shows, if the Internet exponent were the γ = 2.2 ± 0.1 reported in (Pastor-Satorras et al., 2001) rather than γ = 2.5, entropy would account more fully for 60 Chapter 5. Correlated networks and natural disassortativity * * * * * * r r r r r r 0.2 0.2 0.2 0.2 0.2 0.2 0 0 0 0 0 0 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.4 -0.4 -0.4 -0.4 -0.4 -0.4 2 2 2 2 2 2 u d e . d n c i l o b a t e m s r o t c a t e n r e t n i n i e t o r p P 2 P 2.25 2.25 2.25 2.25 2.25 2.25 γ γ γ γ γ γ 2.5 2.5 2.5 2.5 2.5 2.5 Figure 5.4: Level of assortativity that maximizes the entropy, r∗ , for various real-world, scale-free networks, as predicted theoretically by Eq. (5.1) (circles) and as directly measured (horizontal lines), against exponent γ . egalitarian P2P network. Finally, as is typical of social networks, the actor graph is significantly more assortative than predicted, probably due to the ho- mophily mechanism whereby highly connected, big-name actors tend to work together (Newman, 2002, 2003a). 5.4 To sum up... We have shown how the ensemble of networks with a given degree sequence can be partitioned into regions of equally correlated networks and found, using an information-theory approach, that the largest (maximum entropy) region, for the case of scale-free networks, usually displays a certain disassortativity. Therefore, in the absence of knowledge regarding the specific evolutionary forces at work, this should be considered the most likely state. Given the accuracy with which our approach can predict the degree of assortativity of certain empirical networks with no a priori information thereon, we suggest this as a neutral model to decide whether or not particular experimental data require specific mechanisms to account for observed degree-degree correlations. these correlations. Chapter 6 Enhancing robustness to noise via assortativity As we saw in Chapter 4, the performance of attractor neural networks depends crucially on the heterogeneity of the underlying topology’s degree distribution. We take this analysis a step further by examining the effect of degree-degree correlations – or assortativity – on neural-network behaviour. In Chapter 5 we described a method for studying correlated networks and dynamics thereon, both analytically and computationally, which is independent of how the topol- ogy may have evolved. We now make use of this to show how the robustness to noise is greatly enhanced in assortative (positively correlated) neural networks, especially if it is the hub neurons that store the information. 6.1 Background For a dozen years or so now, the study of complex systems has been heavily influenced by results from network science – which one might regard as the fu- sion of graph theory with statistical physics (Newman, 2003c; Boccaletti et al., 2006). Phenomena as diverse as epidemics (Watts and Strogatz, 1998), cel- lular function (Suel et al., 2006), power-grid failures (Buldyrev et al., 2010) or internet routing (Bogun´a et al., 2010), among many others (Arenas et al., 2008a), depend crucially on the structure of the underlying network of in- teractions. One of the earliest systems to have been described as a net- work was the brain, which is made up of a great many neurons connected to each other by synapses (y Ca jal, 1995; Amit, 1989; Abbott and Kepler, 1990; Torres and Varona, 2010). Mathematically, the first neural networks combined the Ising model (Baxter, 1982) with the Hebb learning rule (Hebb, 61 62 Chapter 6. Enhancing robustness to noise via assortativity 1949) to reproduce, very successfully, the storage and retrieval of informa- tion (Amari, 1972; Hopfield, 1982; Amit, 1995). Neurons were simplified to binary variables (like Ising spins) representing firing or non-firing cells. By considering the trivial fully-connected topology, exact solutions could be reached, which at the time seemed more important than attempting to in- troduce biological realism. Subsequent work has tended to focus on consider- ing richer dynamics for the cells rather than on the way in which these are interconnected (Vogels et al., 2005; Torres et al., 2007; Mejias et al., 2010). However, the topology of the brain – whether at the level of neurons and synapses, cortical areas or functional connections – is obviously far from triv- ial (Amaral et al., 2000; Sporns et al., 2004; Egu´ıluz et al., 2005; Arenas et al., 2008b; Bullmore and Sporns, 2009; Johnson et al., 2010a). The number of neighbours a given node in a network has is called its degree, and much attention is paid to degree distributions since they tend to be highly heterogeneous for most real networks. In fact, they are often approximately scale-free (i.e., described by power laws) (Newman, 2003c; Boccaletti et al., 2006; Peretto, 1992; Barab´asi and Oltvai, 2004). By including this topologi- cal feature in a Hopfield-like neural-network model, Torres et al. Torres et al. (2004) found that degree heterogeneity increases the system’s performance at high levels of noise, since the hubs (high degree nodes) are able to retain information at levels well above the usual critical noise. To prove this ana- lytically, the authors considered the configurational ensemble of networks (the set of random networks with a given degree distribution but no degree-degree correlations) and showed that Monte Carlo (MC) simulations were in good agreement with mean-field analysis, despite the approximation inherent to the latter technique when the network is not fully connected. A similar approach can also be used to show how heterogeneity may be advantageous for the per- formance of certain tasks in models with a richer dynamics (Johnson et al., 2008). It is worth mentioning that this influence of the degree distribution on dynamical behaviour is found in many other settings, such as the more general situation of systems of coupled oscillators (Barahona and Pecora, 2002). Another property of empirical networks that is quite ubiquitous is the exis- tence of correlations between the degrees of nodes and those of their neighbours (Pastor-Satorras et al., 2001; Newman, 2002, 2003a). If the average degree- degree correlation is positive the network is said to be assortative, while it is called disassortative if negatively correlated. Most heterogeneous networks are disassortative (Newman, 2003c), which, as described in Chapter 5, seems to be 6.1 Background 63 because this is in some sense their equilibrium (maximum entropy) state given the constraints imposed by the degree distribution (Johnson et al., 2010b). However, there are probably often mechanisms at work which drive systems from equilibrium by inducing different correlations, as appears to be the case for most social networks, in which nodes (people) of a kind tend to group together. This feature, known as assortativity or mixing by degree, is also relevant for processes taking place on networks. For instance, assortative net- works have lower percolation thresholds and are more robust to targeted attack (Newman, 2003a), while disassortative ones make for more stable ecosystems and are – at least according to the usual definition – more synchronizable (Brede and Sinha). The approach usually taken when studying correlated networks computa- tionally is to generate a network from the configuration ensemble and then introduce correlations (positive or negative) by some stochastic rewiring pro- cess (Maslov et al., 2004). A drawback of this method, however, is that results may well then depend on the details of this mechanism: there is no guarantee that one is correctly sampling the phase space of networks with given corre- lations. For analytical work, some kind of hidden variables from which the correlations originate are often considered (Caldarelli et al., 2002; Soderberg, 2002; Bogun´a and Pastor-Satorras, 2003; Fronczak and Fronczak, 2006) – an assumption which can also be used to generate correlated networks compu- tationally (Bogun´a and Pastor-Satorras, 2003). This can be a very powerful method for solving specific network models. However, it may not be appropri- ate if one wishes to consider all possible networks with given degree-degree cor- relations, independently of how these may have arisen. In this chapter, we get round the problem by making use of the method put forward by Johnson et al. (2010b) (and described in Chapter 5) whereby the ensemble of all networks with given correlations can be considered theoretically without recurring to hidden variables (de Franciscis et al., 2011). Furthermore, we show how this approach can be used computationally to generate random networks that are representative of the ensemble of interest (i.e., they are model-independent). In this way, we study the effect of correlations on a simple neural network model and find that assortativity increases performance in the face of noise – partic- ularly if it is the hubs that are mainly responsible for storing information (and it is worth mentioning that there is experimental evidence suggestive of a main functional role played by hub neurons in the brain (Morgan and Soltesz, 2008; Bonifazi et al., 2009)). The good agreement between the mean-field analysis 64 Chapter 6. Enhancing robustness to noise via assortativity and our MC simulations bears witness both to the robustness of the results as regards neural systems, and to the viability of using this method for studying dynamics on correlated networks. 6.2 Preliminary considerations 6.2.1 Model neurons on networks The attractor neural network model put forward by Hopfield (Hopfield, 1982) consists of N binary neurons, each with an activity given by the dynamic vari- able si = ±1. Every time step (MCS), each neuron is updated according to the stochastic transition probability P (si → ±1) = 1 2 [1 ± tanh (hi /T )] (paral- lel dynamics), where the field hi is the combined effect on i of all its neighbours, hi = Pj wij sj , and T is a noise parameter we shall call temperature, but which represents any kind of random fluctuations in the environment. This is the same as the Ising model for magnetic systems, and the transition rule can be derived from a simple interaction energy such that aligned variables s (spins) contribute less energy than if they were to take opposite values. However, this system can store P given configurations (memory patterns) ξ ν i = ±1 by having the interaction strengths (synaptic weights) set according to the Hebb rule (Hebb, 1949): wij ∝ PP ν=1 ξ ν i ξ ν j . In this way, each pattern becomes an attractor of the dynamics, and the system will evolve towards whichever one is closest to the initial state it is placed in. This mechanism is called associative memory, and is nowadays used routinely for tasks such as image identifica- tion. What is more, it has been established that something similar to the Hebb rule is implemented in nature via the processes of long-term potenti- ation and depression at the synapses (Malenka and Nicoll, 1999; Roo et al., 2008; Rodr´ıguez-Moreno and Paulsen, 2008; Kwag and Paulsen, 2009), and this phenomenon is indeed required for learning (Gruart et al., 2006). To take into account the topology of the network, we shall consider the weights to be of the form wij = ωij aij , where the element aij of the adjacency matrix represents the number of directed edges (usually interpreted as synapses in a neural network) from node j to node i, while ω stores the patterns, as before: P 1 Xν=1 hki For the sake of coherence with previous work, we shall assume a to be sym- metric (i.e., the network is undirected), so each node is characterized by a ωij = i ξ ν ξ ν j . 6.2 Preliminary considerations 65 single degree ki = Pj aij . However, all results are easily extended to directed networks – in which nodes have both an in degree, k in i = Pj aij , and an out degree, kout = Pj aj i – by bearing in mind it is only a neuron’s pre-synaptic i neighbours that influence its behaviour. The mean degree of the network is hki, where the angles stand for an average over nodes1 : h·i ≡ N −1 Pi (·). 6.2.2 Network ensembles When one wishes to consider a set of networks which are randomly wired while respecting certain constraints – that is, an ensemble – it is usually useful to define the expected value of the adjacency matrix2 , E (a) ≡ ǫ. The element ǫij of this matrix is the mean value of aij obtained by averaging over the en- semble. For instance, in the Erdos-R´enyi (ER) ensemble all elements (outside the diagonal) take the value ǫER ij = hki/N , which is the probability that a given pair of nodes be connected by an edge. For studying networks with a given degree sequence, (k1 , ...kN ), it is common to assume the configuration ensemble, defined as ǫconf ij = kikj hkiN This expression can usually be applied also when the constraint is a given de- gree distribution, p(k), by integrating over p(ki ) and p(kj ) where appropriate. One way of deriving ǫconf is to assume one has ki dangling half-edges at each node i; we then randomly choose pairs of half-edges and join them together until the network is wired up. Each time we do this, the probability that we join i to j is kikj /(hkiN )2 , and we must perform the operation hkiN times. Bianconi showed that this is also the solution for Barab´asi-Albert evolved net- works (Bianconi, 2002). However, we should bear in mind that this result is only strictly valid for networks constructed in certain particular ways, such as in these examples. It is often implicitly assumed that were we to average over all random networks with a given degree distribution, the mean adjacency matrix obtained would be ǫconf . However, as we discussed in Chapter 5, this is not in fact necessarily true (Johnson et al., 2010b). 1 In directed networks the mean in degree and the mean out degree necessarily coincide, whatever the forms of the in and out distributions. 2As in statistical physics, one can consider the microcanonical ensemble, in which each element (network) satisfies the constraints exactly, or the canonical ensemble, where the constraints are satisfied on average (Bianconi, 2009). Throughout this work, we shall refer to canonical ensembles. 66 Chapter 6. Enhancing robustness to noise via assortativity 102 > k < - n n k 101 100 10-2 10-4 ) k ( P β=-0.5 β=0 β=0.5 k-2.5 101 102 103 k0.5 101 k-0.5 102 k Figure 6.1: Mean-nearest-neighbour functions knn (k) for scale-free networks with β = −0.5 (disassortative), 0.0 (neutral), and 0.5 assortative, generated according to the algorithm described in Sec. 6.3.2. Inset: degree distribution (the same in all three cases). Other parameters are γ = 2.5, hki = 12.5, N = 104 . 6.2.3 Correlated networks In the configuration ensemble, the expected value of the mean degree of the i Pj ǫconf neighbours of a given node is knn,i = k−1 kj = hk2i/hki, which is in- ij dependent of ki . However, as mentioned above, real networks often display degree-degree correlations, with the result that knn,i = knn(ki ). If knn (k) in- creases with k , the network is said to be assortative – whereas it is disassorta- tive if it decreases with k (see Fig. 6.1). This is from the more general nomen- clature (borrowed form sociology) in which sets are assortative if elements of a kind group together, or assort. In the case of degree-degree correlated net- works, positive assortativity means that edges are more than randomly likely to occur between nodes of a similar degree. The ensemble of all networks with a given degree sequence (k1 , ...kN ) con- tains a subset for all members of which knn (k) is constant (the configuration ensemble), but also subsets displaying other functions knn (k). We can iden- tify each one of these subsets (regions of phase space) with an expected ad- jacency matrix ǫ which simultaneously satisfies the following conditions: i) Pj kj ǫij = kiknn (ki ), ∀i (by definition of knn(k)), and ii) Pj ǫij = ki , ∀i (for consistency). As we showed in Chapter 5, the general solution to this problem 6.2 Preliminary considerations 67 is a matrix of the form ǫij = N (cid:20) (kikj )ν j + hkν i(cid:21) , kikj f (ν ) + Z dν i − kν hkν i − kν hkiN where ν ∈ R and the function f (ν ) is determined by knn (k) (Johnson et al., 2010b). (If the network were directed, then ki = k in and kj = kout in this i j expression.) This yields (6.1) (6.2) + Z dν f (ν )σν+1 (cid:20) kν−1 knn (k) = hk2i k (cid:21) 1 hkν i − hki (the first term being the result for the configuration ensemble), where σb+1 ≡ hk b+1i − hkihk bi. This means that ǫ is not just one possible way of obtaining correlations according to knn (k); rather, there is a two-way mapping between ǫ and knn (k): every network with this particular function knn (k) and no other ones are contained in the ensemble defined by ǫ. Thanks to this, if we are able to consider random networks drawn according to this matrix (whether we do this analytically or computationally; see Section 6.3.2), we can be con- fident that we are correctly taking account of the whole ensemble of interest. In other words, whatever the reasons behind the existence of degree-degree correlations in a given network, we can study the effects of these with only information on p(k) and knn (k) by obtaining the associated matrix ǫ. This is not to say, of course, that all topological properties are captured in this way: a particular network may have other features – such as higher order correlations, modularity, etc. – the consideration of which would require concentrating on a sub-partition of those with the same p(k) and knn (k). But this is not our purpose here. In many empirical networks, knn (k) has the form knn (k) = A + Bkβ , with A, B > 0 (Boccaletti et al., 2006; Pastor-Satorras et al., 2001) – the mixing being assortative if β is positive, and disassortative when negative. Such a case is fitted by Eq. (6.2) if δ (ν − β − 1) − δ (ν − 1)(cid:21) , f (ν ) = C (cid:20) σ2 σβ+2 with C a positive constant, since this choice yields + C σ2 (cid:20) kβ hkβ+1i − knn(k) = hk2i hki hki (cid:21) . 1 (6.3) (6.4) 68 Chapter 6. Enhancing robustness to noise via assortativity In Chapter 5 we discussed how the most likely configurations for networks with scale-free degree distributions (p(k) ∼ k−γ ) and correlations given by Eq. (6.4) are generally disassortative. We also showed that the maximum entropy is usually obtained for values of C close to one. Here, we shall use this result to justify concentrating on correlated networks with C = 1, so that the only parameter we need to take into account is β . It is worth mentioning that Pastor-Satorras et al. originally suggested using this exponent as a way of quantifying correlations (Pastor-Satorras et al., 2001), since this seems to be the most relevant magnitude. Because β does not depend directly on p(k) (as r does), and can be defined for networks of any size (whereas r , in very heterogeneous networks, always goes to zero for large N due to its normaliza- tion (Dorogovtsev et al., 2005)), we shall henceforth use β as our assortativity parameter. So, after plugging Eq. (6.3) into Eq. (6.1), we find that the ensemble of networks exhibiting correlations given by Eq. (6.4) (and C = 1) is defined by the mean adjacency matrix 1 N σ2 σβ+2 ǫij = [ki + kj − hki] N (cid:20) (kikj )β+1 j + hkβ+1i(cid:21) . 1 hkβ+1i − kβ+1 i − kβ+1 6.3 Analysis and results + (6.5) 6.3.1 Mean field Let us consider the single-pattern case (P = 1, ξi = ξ 1 i ). Substituting the adjacency matrix a for its expected value ǫ (as given by Eq. (6.5)) in the ex- pression for the local field at i – which amounts to a mean-field approximation – we have hi = 1 hki )(cid:21) µ0 ξi (cid:26)(cid:20)(ki − hki) + σ2 (hkβ+1i − kβ+1 i σβ+2 i − hkβ+1i)µβ+1(cid:27) , σ2 (kβ + hkiµ1 + σβ+2 µα ≡ hkα i ξisi i hkαi for α = 0, 1, β + 1. These order parameters measure the extent to which the system is able to recall information in spite of noise (Johnson et al., 2008). where we have defined 6.3 Analysis and results 69 For the first order we have µ0 = m ≡ hξisi i, the standard overlap measure in neural networks (analogous to magnetization in magnetic systems), which takes account of memory performance. However, µ1 , for instance, weighs the sum with the degree of each node, with the result that it measures information per synapse instead of per neuron. Although the overlap m is often assumed to represent, in some sense, the mean firing rate of neurological experiments, it is possible that µ1 is more closely related to the empirical measure, since the total electric potential in an area of tissue is likely to depend on the number of synapses transmitting action potentials. In any case, a comparison between the two order parameters is a good way of assessing to what extent the performance of neurons depends on their degree – larger-degree model neurons can in general store information at higher temperatures than ones with smaller degree can (Torres et al., 2004). Substituting si for its expected value according to the transition probability, si → tanh(hi/T ), we have, for any α, i ξisi i = hkα hkα i ξi tanh(hi /T )i; or, equivalently, the following 3-D map of closed coupled equations for the macroscopic overlap observables µ0 , µ1 and µβ+1 – which describes, in this mean-field approximation, the dynamics of the system: µ0(t + 1) = Z p(k) tanh[F (t)/(hkiT )]dk 1 hki Z p(k)k tanh[F (t)/(hkiT )]dk µ1(t + 1) = 1 hkβ+1i Z p(k)kβ+1 tanh[F (t)/(hkiT )]dk , µβ+1(t + 1) = (6.6) with F (t) ≡ (kµ0(t) + hkiµ1(t) − hkiµ0(t)) + σ2 σβ+2 [kβ+1 (µβ+1(t) − µ0(t)) + hkβ+1i(µ0(t) − µβ+1(t))]. This can be easily computed for any degree distribution p(k). Note that taking β = 0 (the uncorrelated case) the system collapses to the 2-D map obtained by 70 Chapter 6. Enhancing robustness to noise via assortativity Torres et al. (2004), while it becomes the typical 1-D case for a homogeneous p(k) – say a fully-connected network (Hopfield, 1982). It is in principle pos- sible to do similar mean-field analysis for any number P of patterns, but the map would then be 3P -dimensional, making the problem substantially more complex. At a critical temperature Tc , the system will undergo the characteristic second order phase transition from a phase in which it exhibits memory (akin to ferromagnetism) to one in which it does not (paramagnetism). To obtain this critical temperature, we can expand the hyperbolic tangent in Eqs. (6.6) around the trivial solution (µ0 , µ1 , µβ+1) ≃ (0, 0, 0) and, keeping only linear terms, write µ0 = µ1/Tc , µ1 = µβ+1 = 1 hki2Tc (cid:2)hki2µ1 + σ2µβ+1(cid:3) , 1 Tchkihkβ+1i hσβ+2µ0 σ2 σβ+2 (cid:0)hkβ+1 i2 − hk2(β+1) i(cid:1) µ0 σβ+2 (cid:0)hkβ+1i2 − hk2(β+1) i(cid:1) µβ+1(cid:21) . σ2 + hkihkβ+1iµ1 − + Defining A ≡ B ≡ σ2 hki2 , σ2 σβ+2 hk2(β+1) i − hkβ+1 i2 hkihkβ+1i , σβ+2 hkihkβ+1i Tc will be the solution to the third order polynomial equation: D ≡ , c − (B + 1)T 2 T 3 c + (B − A)Tc + A(B − D) = 0. (6.7) Note that for neutral (i.e., uncorrelated) networks, β = 0, and so A = B = D . We then have Tc = hk2i/hki2 , as expected (Johnson et al., 2008). 6.3 Analysis and results 71 6.3.2 Generating correlated networks Given a degree distribution p(k), the ensemble of networks compatible with this constraint and with degree-degree correlations according to Eq. (6.4) (with some exponent β ) is defined by the mean adjacency matrix ǫ of Eq. (6.5) – as described in Section 6.2.3 and by Johnson et al. (2010b). Therefore, although there will generally be an enormous number of possible networks in this volume of phase space, we can sample them correctly simply by generating them according to ǫ. To do this, first we have to assign to each node a degree drawn from p(k). If the elements of ǫ were probabilities, it would suffice then to connect each pair of nodes (i, j ) with probability ǫij to generate a valid network. Strictly speaking, ǫ is an expected value, which in certain cases can be greater than one. To get round this, we write a probability matrix p = ǫ/a with a some value such that all elements of p are smaller than one. If we then take random pairs of nodes (i, j ) and, with probability pij , place an edge between them, repeating the operation until 1 2 hkiN edges have been placed, the expected value of edges joining i and j will be ǫij . This method is like the hidden variable technique (Bogun´a and Pastor-Satorras, 2003) in that edges are placed with a predefined probability (which is why the resulting ensemble is canonical). The difference lies in the fact that in the method here described correlations only depend on the degrees of nodes. We are interested here in neural networks, in which a given pair of nodes can be joined by several synapses, so we shall not impose the restriction of so-called simple networks of allowing only one edge at most per pair. We shall, however, consider networks with a structural cutoff: ki < phkiN , ∀i (Bianconi, 2008). This ensures that, at least for β ≤ 0, all elements of ǫ are indeed smaller than one. Because we can expect effects due to degree-degree correlations to be largest when p(k) is very broad, and since most networks in nature and technol- ogy seem to exhibit approximately power-law degree distributions (Newman, 2003c; Arenas et al., 2008a; Peretto, 1992; Barab´asi and Oltvai, 2004), we shall here test our general theoretical results against simulations of scale-free net- works: p(k) ∼ k−γ . This means that a network (or the region of phase space to which it belongs) is characterized by the set of parameters {hki, N , γ , β }. 72 Chapter 6. Enhancing robustness to noise via assortativity 1 0.5 0 0 N=104 N=3x104 N=5x104 β=0.5 4 8 12 1 0.5 1 µ β=-0.5 β=0 β=0.5 0 0 1 2 3 T 4 5 6 7 Figure 6.2: Stable stationary value of the weighted overlap µ1 against tem- perature T for scale-free networks with correlations according to knn ∼ kβ , for β = −0.5 (disassortative), 0.0 (neutral), and 0.5 (assortative). Symbols from MC simulations, with errorbars representing standard deviations, and lines from Eqs. (6.6). Other network parameters as in Fig. 6.1. Inset: µ1 against T for the assortative case (β = 0.5) and different system sizes: N = 104 , 3 · 104 and 5 · 104 . 6.3.3 Assortativity and dynamics In Fig. 6.2 we plot the stationary value of µ1 against the temperature T , as obtained from simulations and Eqs. (6.6), for disassortative, neutral and assortative networks. The three curves are similar at low temperatures, but as T increases their behaviour becomes quite different. The disassortative network is the least robust to noise. However, the assortative one is capable of retaining some information at temperatures considerably higher than the critical value, Tc = hk2i/hki, of neutral networks. A comparison between µ1 and µ0 (see Fig. 6.3) shows that it is the high degree nodes that are mainly responsible for this difference in performance. This can be seen more clearly in Fig. 6.4, which displays the difference µ1 − µ0 against T for the same networks. It seems that, because in an assortative network a sub-graph of hubs will have more edges than in a disassortative one, it has a higher effective critical temperature. Therefore, even when most of the nodes are acting randomly, the set of nodes of sufficiently high degree nevertheless displays associative memory. 6.3 Analysis and results 73 1 0.5 x µ µ 0µ 1µβ+1 β=0.5 0 0 1 2 3 4 5 6 7 T Figure 6.3: Stable stationary values of order parameters µ0 , µ1 and µβ+1 against temperature T , for assortative networks according to β = 0.5. Symbols from MC simulations, with errorbars representing standard deviations, and lines from Eqs. (6.6). Other parameters as in Fig. 6.1. 0.3 0 µ 0.15 - 1 µ β=-0.5 β=0 β=0.5 2N-1/2 0 0 4 T 8 10 Figure 6.4: Difference between the stationary values µ1 and µ0 for networks with β = −0.5 (disassortative), 0.0 (neutral) and 0.5 (assortative), against temperature. Symbols from MC simulations, with errorbars representing stan- dard deviations, and lines from Eqs. (6.6). Line shows the expected level of fluctuations due to noise, ∼ N − 1 2 . Other parameters as in Fig. 6.1. 74 Chapter 6. Enhancing robustness to noise via assortativity The phase diagram if Fig. 6.5 shows the critical temperature, Tc , as ob- tained from Eq. (6.7). In addition to the effect reported by Torres et al. (2004) whereby the Tc of scale-free networks grows with degree heterogeneity (decreasing γ ), it also increases very significantly with positive degree-degree correlations (increasing β ). At large values of N , the critical temperature scales as Tc ∼ N b , with b ≥ 0 a constant. However, because the moments of k appearing in the coefficients of Eq. (6.7) can have different asymptotic behaviour depending on the values of γ and β , the scaling exponent b differs from one region to another in the space of these parameters. These are the seven regions shown in Fig. 6.6, along with the scaling behaviour exhibited by each one. This can be seen explicitely in Fig. 6.7, where Tc , as obtained from MC simulations, is plotted against N for cases in each of the regions with γ < 3. In each case, the scaling is as given by Eq. (6.7) and shown in Fig. 6.6. For the four regions with γ < 3, from lowest to highest assortativity we have scaling exponents which are dependent on: only γ (region I), only β (region II), both γ and β (region III), and, perhaps most interestingly, neither of the two (region IV) – with Tc scaling, in the latter case, as √N . As for the more homogeneous γ > 3 part, regions V and VI have a diverging critical temperature despite the fact that the second moment of p(k) is finite, simply as a result of assortativity. The case in which more than one pattern are stored (P > 1) can be explored numerically. Assuming there are P uncorrelated patterns, we have an order parameter µν 1 for each pattern ν . A global measure of the degree to which there is memory can be captured by the parameter ζ , where ζ 2 ≡ 1 1 + P /N (µν 1 )2 . P Xν=1 Notice that the normalization factor is due to the fact that if one pattern is condensed – i.e., µ1 . 1 – the others have µν ∼ 1/√N , ν = 2, ..P , and so ζ ≃ 1. Figure 6.8 shows how ζ decreases with T in variously correlated networks for P = 3 (left panel) and P = 10 patterns (right panel). The behaviour is not qualitatively different from that observed for the single-pattern case in the main panel of Fig. 6.2, suggesting that the influence of assortativity we report is robust as to the number of patterns stored, P . 6.3 Analysis and results 75 P γ=2.5 8 6 γ=3 4 T γ=3.5 2 F 0.5 0 1 -1 -0.5 0 β Figure 6.5: Phase diagrams for scale-free networks with γ = 2.5, 3, and 3.5. Lines show the critical temperature Tc marking the second-order transition from a memory (ferromagnetic) phase to a memoryless (paramagnetic) one, against the assortativity β , as given by Eq. (6.7). Other parameters as in Fig. 6.1. β+2 β+3 2β+3 4 3.5 (0) finite Tc ∝ Tc N(3+2β-γ)/2 (V) β/2 ∝ N (VI)Tc γ 3 ∝ N(3-γ)/4 (I) Tc 2.5 2 -1 (II) ∝ N-β/2 Tc -0.5 ∝ N(3+β-γ)/2 (III)Tc ∝ N1/2 (IV)Tc 0.5 1 0 β Figure 6.6: Parameter space β − γ partitioned into the regions in which b(β , γ ) has the same functional form – where b is the scaling exponent of the critical temperature: Tc ∼ N b . Exponents obtained by taking the large N limit in Eq. (6.7). 76 Chapter 6. Enhancing robustness to noise via assortativity β=-0.8 β=-0.35 β=0 β=0.9 N(-γ+3)/4 N(-β)/2 N(-γ+β+3)/2 N1/2 10 c T 1 102 103 104 105 N Figure 6.7: Examples of how Tc scales with N for networks belonging to regions I, II, III and IV of Fig. 6.6 (β = −0.8, −0.35, 0.0 and 0.9, respec- tively). Symbols from MC simulations, with errorbars representing standard deviations, and slopes from Eq. (6.7). All parameters – except for β and N – are as in Fig. 6.1. 6.4 Discussion We have shown that assortative networks of simple model neurons are able to exhibit associative memory in the presence of levels of noise such that uncor- related (or disassortative) networks cannot. This may appear to be in con- tradiction with a recent result obtained using spectral graph analysis – that synchronizability of a set of coupled oscillators is highest for disassortative networks (Brede and Sinha). A synchronous state of model oscillators and a memory phase of model neurons are both sets of many simple dynamical el- ements coupled via a network in such a way that a macroscopically coherent situation is maintained (Barahona and Pecora, 2002). Obviously both systems require the effective transmission of information among the elements. So why are opposite results as regards the influence of topology reported for each sys- tem? The answer is simple: whereas the definition of a synchronous state is that every single element oscillate at the same frequency, it is precisely when most elements are actually behaving randomly that the advantages to assor- tativity we report become apparent. In fact, it can be seen in Fig. 6.2 that at low temperatures disassortative networks perform the best, although the ef- 6.4 Discussion 77 1 2 / 1 ) 2 ) 0.5 1 ν µ ( ν Σ ( 0 0 β=-0.5β=0 β=0.5 4 T 8 β=-0.5β=0 β=0.5 0 4 T 8 Figure 6.8: Global order parameter ζ for assortative (β = 0.5), neutral (β = 0.0) and disassortative (β = −0.5) networks with P = 3 (left panel) and P = 10 (right panel) stored patterns. Symbols from MC simulations, with errorbars representing standard deviations. All parameters are as in Fig. 6.1. fect is small. This is reminiscent of percolation: at high densities of edges the giant component is larger in disassortative networks, but in assortative ones a non-vanishing fraction of nodes remain interconnected even at densities below the usual percolation threshold (Newman, 2002, 2003a). Because in the case of targeted attacks it is this threshold which is taken as a measure of resilience, we say that assortative networks perform the best. The relevance of partial synchronization and the important role of hubs have already been noted for systems of (weakly) coupled oscillators (G´omez-Gardenes et al., 2007; Pereira, 2010) – for which, however, assortativity has not been expected to be of con- sequence (Pereira, 2010). In general, the optimal network for good conditions (i.e., complete synchronization, high density of edges, low levels of noise) is not necessarily the one which performs the best in bad conditions (partial synchro- nization, low density of edges, high levels of noise). It seems that optimality – whether in resilience or robustness – should thus be defined for particular conditions. We have used the technique suggested by Johnson et al. (2010b) to study the effect of correlations on networks of model neurons, but many other sys- 78 Chapter 6. Enhancing robustness to noise via assortativity tems of dynamical elements should be susceptible to a similar treatment. In fact, Ising spins (Bianconi, 2002), Voter Model agents (Suchecki et al., 2005), or Boolean nodes (Peixoto, 2010), for instance, are similar enough to binary neurons that we should expect similar results for these models. If a moral can be drawn, it is that persistence of partial synchrony, or coherence of a subset of highly connected dynamical elements, can sometimes be as relevant (or more so) as the possibility of every element behaving in the same way. In the case of real brain cells, experiments suggest that hub neurons play key functional roles (Morgan and Soltesz, 2008; Bonifazi et al., 2009). From this point of view, there may be a selective pressure for brain networks to become assortative – although, admittedly, this organ engages in such complex behaviour that there must be many more functional constraints on its structure than just a high robustness to noise. Nevertheless, it would be interesting to investigate this as- pect of biological systems experimentally. For this, it should be borne in mind that heterogeneous networks have a natural tendency to become disassortative, so it is against the expected value of correlations discussed by Johnson et al. (2010b) that empirical data should be contrasted in order to look for meaning- ful deviations towards assortativity. Similarly, it may be necessary to take into account the correlations that could emerge due to the spatial layout of neurons (Kaiser et al., 2007; Johnson et al., 2011). In any case, it would be in areas of the cortex specifically related to memory – such as the temporal (long-term memory) (Miyashita, 1988; Sakai and Miyashita, 1991) or prefrontal (short- term memory) (Camperi and Wang, 1998b; Compte et al., 2003) lobes – that this effect might be relevant. A curious fact that would seem to support our hy- pothesis is that whereas the vast ma jority of non-social networks are disassor- tative (Newman, 2003c), one that appears actually to be strongly assortative is the functional network of the human cortex (Egu´ıluz et al., 2005). Chapter 7 Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning Short-term memory cannot in general be explained the way long-term memory can – as a gradual modification of synaptic conductances – since it takes place too quickly. Theories based on some form of cellular bistability, however, do not seem to be able to account for the fact that noisy neurons can collectively store information in a robust manner. We show how a sufficiently clustered network of simple model neurons can be instantly induced into metastable states capable of retaining information for a short time. Cluster Reverbera- tion, as we call it, could constitute a viable mechanism available to the brain for robust short-term memory with no need of synaptic learning. Relevant phe- nomena described by neurobiology and psychology, such as power-law statistics of forgetting avalanches, emerge naturally from this mechanism. 7.1 Slow but sure, or fast and fleeting? Of all brain phenomena, memory is probably one of the best understood (Amit, 1989; Abbott and Kepler, 1990; Torres and Varona, 2010). Consider a set of many neurons, defined as elements with two possible states (firing or not firing, one or zero) connected among each other in some way by synapses which carry a proportion of the current let off by a firing neuron to its neighbours; the probability that a given neuron has of firing at a certain time is then some 79 80 Chapter 7. Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning function of the total current it has just received. Such a simplified model of the brain is able to store and retrieve information, in the form of patterns of activity (i.e., particular configurations of firing and non-firing neurons) when the synaptic conductances, or weights, have been appropriately set according to a learning rule (Hebb, 1949). Because each of the stored patterns becomes an attractor of the dynamics, the system will evolve towards whichever of the patterns most resembles the initial configuration. Artificial systems used for tasks such as pattern recognition and classification, as well as more realistic neural network models that take into account a variety of subcellular processes, all tend to rely on this basic mechanism, known as Associative Memory (Amari, 1972; Hopfield, 1982). Synaptic conductances in animal brains have indeed been found to be- come strengthened or weakened during learning, via the biochemical processes of long-term potentiation (LTP) and depression (LTD) (Malenka and Nicoll, 1999; Gruart et al., 2006; Roo et al., 2008; Rodr´ıguez-Moreno and Paulsen, 2008; Kwag and Paulsen, 2009). Further support for the hypothesis that such a mechanism underlies long-term memory (LTM) comes from psychology, where it is being found more and more that so-called connectionist models fit in well with observed brain phenomena (Marcus and G.F., 2001; Frank, 1997). How- ever, some memory processes take place on timescales of seconds or less and in many instances cannot be accounted for by LTP and LTD (Durstewitz et al., 2000), since these require at least minutes to be effected (Lee et al., 1980; Klintsova and Greenough, 1999). For example, Sperling found that visual stimuli are recalled in great detail for up to about one second after expo- sure (iconic memory) (Sperling, 1960); similarly, acoustic information seems to linger for three or four seconds (echoic memory) (Cowan, 1984). In fact, it appears that the brain actually holds and continually updates a kind of buffer in which sensory information regarding its surroundings is maintained (sensory memory) (Baddeley, 1999). This is easily observed by simply closing one’s eyes and recalling what was last seen, or thinking about a sound after it has finished. Another instance is the capability referred to as working mem- ory (Durstewitz et al., 2000; Baddeley and A.D., 2003): just as a computer requires RAM for its calculations despite having a hard drive for long term storage, the brain must continually store and delete information to perform almost any cognitive task. To some extent, working memory could consist in somehow labelling or bringing forward previously stored concepts, like when one is asked to remember a particular sequence of digits or familiar shapes. 7.1 Slow but sure, or fast and fleeting? 81 But we are also able to manipulate – if perhaps not quite so well – shapes and symbols we have only just become acquainted with, too recently for them to have been learned synaptically. We shall here use short-term memory (STM) to describe the brain’s ability to store information on a timescale of seconds or less1 . Evidence that short-term memory is related to sensory information while long-term memory is more conceptual can again be found in psychology. For instance, a sequence of similar sounding letters is more difficult to retain for a short time than one of phonetically distinct ones, while this has no bear- ing on long-term memory, for which semantics seems to play the main role (Conrad, 1964a,b); and the way many of us think about certain concepts, such as chess, geometry or music, is apparently quite sensorial: we imagine positions, surfaces or notes as they would look or sound. Most theories of short-term memory – which almost always focus on working memory – make use of some form of previously stored information (i.e., of synaptic learning) and so can account for the labelling tasks referred to above but not for the instant recall of novel information (Wang, 2001; Barak and Tsodyks, 2007; Roudi and Latham; Mongillo et al., 2008; Mejias and Torres, 2009). Attempts to deal with the latter have been made by proposing mechanisms of cel lular bistability : neurons are assumed to retain the state they are placed in (such as firing or not firing) for some period of time thereafter (Camperi and Wang, 1998a; Teramae and Fukai, 2005; Tarnow, 2008). Although there may indeed be subcellular processes leading to a certain bistability, the main problem with short-term memory depending exclusively on such a mechanism is that if each neuron must act independently of the rest the patterns will not be robust to random fluctuations (Durstewitz et al., 2000) – and the behaviour of individual neurons is known to be quite noisy (Compte et al., 2003). It is worth pointing out that one of the strengths of Associative Memory is that the behaviour of a given neuron depends on many neighbours and not just on itself, which means that robust global recall can emerge despite random fluctuations at an 1We should mention that sensory memory is usually considered distinct from STM – and probably has a different origin – but we shall use “short-term memory” generically since the mechanism we propose in this paper could be relevant for either or both phenomena. On the other hand, the recent flurry of research in psychology and neuroscience on working memory has lead to this term sometimes being used to mean short-term memory; strictly speaking, however, working memory is generally considered to be an aspect of cognition which operates on information stored in STM. 82 Chapter 7. Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning individual level. Something that, at least until recently, most neural network models have failed to take into account is the structure of the network – its topology – it often being assumed that synapses are placed among the neurons completely at random, or even that all neurons are connected to all the rest (a mathe- matically convenient but unrealistic situation). Although relatively little is yet known about the architecture of the brain at the level of neurons and synapses, experiments have shown that it is heterogeneous (some neurons have very many more synapses than others), clustered (two neurons have a higher chance of be- ing connected if they share neighbours than if not) and highly modular (there are groups, or modules, with neurons forming synapses preferentially to those in the same module) (Sporns et al., 2004; Johnson et al., 2010a). This chapter describes the main result of Ref. (Johnson et al., 2011) – namely, that it suf- fices to use a more realistic topology, in particular one which is modular and/or clustered, for a randomly chosen pattern of activity the system is placed in to be metastable. This means that novel information can be instantly stored and retained for a short period of time in the absence of both synaptic learning and cellular bistability. The only requisite is that the patterns be coarse grained versions of the usual patterns – that is, whereas it is often assumed that each neuron in some way represents one bit of information, we shall allocate a bit to a small group or neurons2 (four or five can be enough). If The mechanism, which we call Cluster Reverberation, is very simple. neurons in a group are more highly connected to each other than to the rest of the network, either because they form a module or because the network is significantly clustered, they will tend to retain the activity of the group: when they are all initially firing, they each continue to receive many action potentials and so go on firing, whereas if they start off silent, there is not usually enough input current from the outside to set them off. The fact that each neuron’s state depends on its neighbours conferres to the mechanism a certain robustness in the face of random fluctuations. This robustness is par- ticularly important for biological neurons, which as mentioned are quite noisy. Furthermore, not only does the limited duration of short-term memory states emerge naturally from this mechanism (even in the absence of interference from new stimuli) but this natural forgetting follows power-law statistics, as 2This does not, of course, mean that memories are expected to be encoded as bitmaps. Just as with individual neurons, positions or orientations, say, could be represented by the activation of particular sets of clusters. 7.2 The simplest neurons on modular networks 83 in experimental settings (Wixted and Ebbesen, 1991, 1997; Sikstrom, 2002). The process is reminiscent both of block attractors in ordinary neural net- works (Dominguez et al., 2009) and of domains in magnetic materials (A. and R., 1998), while Munoz et al. have recently highlighted a similarity with Griffiths phases on networks (Munoz et al., 2010). It can also be interpreted as a multi- scale phenomenon: the mesoscopic clusters take on the role usually played by individual neurons, yet make use of network properties. Although the mecha- nism could also work in conjunction with other ones, such as synaptic learning or cellular bistability, we shall illustrate it by considering the simplest model which has the necessary ingredients: a set of binary neurons linked by synapses of uniform weight according to a topology whose modularity or clustering we shall tune. As with Associative Memory, this mechanism of Cluster Reverber- ation appears to be simple and robust enough not to be qualitatively affected by the complex subcellular processes incorporated into more realistic neuron models – such as integrate-and-fire or Hodgkin-Huxley neurons. However, such refinements are probably needed to achieve graded persistent activity, since the mean frequency of each cluster could then be set to a certain value. 7.2 The simplest neurons on modular networks We consider a network of N model neurons, with activities si = ±1. The topol- ogy is given by the adjacency matrix aij = {1, 0}, each element representing the existence or absence of a synapse from neuron j to neuron i (a need not be symmetric). In this kind of model, each edge usually has a synaptic weight associated, ωij ∈ R; however, we shall here consider these to have all the same value: ωij = ω ∀i, j . Neurons are updated in parallel (Little dynamics) at each time step, according to the stochastic transition rule T (cid:19) + tanh (cid:18) hi P (si → ±1) = ± 1 2 1 2 , where the field of neuron i is defined as aij sj hi = ω N Xj and T is a parameter we shall call temperature. First of all, we shall consider the network defined by a to be made up of M distinct modules. To achieve this, we can first construct M separate random directed networks, each with n = N/M nodes and mean degree (mean 84 Chapter 7. Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning number of neighbours) hki. Then we evaluate each edge and, with probability λ, eliminate it, to be substituted for another edge between the original post- synaptic neuron and a new pre-synaptic neuron chosen at random from among any of those in other modules3 . Note that this protocol does not alter the number of pre-synaptic neighbours of each node, k in i = Pj aij (although the number of post-synaptic neurons, kout i = Pj aj i , can vary). The parameter λ can be seen as a measure of modularity of the partition considered, since it coincides with the expected value of the proportion of edges that link different modules. In particular, λ = 0 defines a network of disconnected modules, while λ = 1 − M −1 yields a random network in which this partition has no modularity. If λ ∈ (1 − M −1 , 1), the partition is less than randomly modular – i.e., it is quasi-multipartite (or multipartite if λ = 1). If the size of the modules is of the order of hki, the network will also be highly clustered. Taking into account that the network is directed, let us define the clustering coefficient Ci as the probability, given that there is a synapse from neuron i to a neuron j and from another neuron l to i, that there be a synapse from j to l: that is, that there exist a feedback loop i → j → l → i. Then, assuming M ≫ 1, the expected value of the clustering coefficient C ≡ hCii is C & hki − 1 n − 1 (1 − λ)3 . 7.3 Cluster Reverberation A memory pattern, in the form of a given configuration of activities, {ξi = ±1}, can be stored in this system with no need of prior learning. Imagine a pattern such that the activities of all n neurons found in any module are the same, i.e., ξi = ξµ(i) , where the index µ(i) denotes the module that neuron i belongs to. This can be thought of as a coarse graining of the standard idea of memory patterns, in which each neuron represents one bit of information. In our scheme, each module represents – and stores – one bit. The system can be induced into this configuration via the application of an appropriate stimulus (see Fig. 7.1): the field of each neuron will be altered for just one time step according to hi → hi + δξµ(i) , ∀i, 3We do not allow self-edges (although these can occur in reality) since they can be regarded as a form of cellular bistability. 7.3 Cluster Reverberation 85 Figure 7.1: Diagram of a modular network composed of four five-neuron clus- ters. The four circles enclosed by the dashed line represent the stimulus: each is connected to a particular module, which adopts the input state (red or blue) and retains it after the stimulus has disappeared via Cluster Reverberation. where the factor δ is the intensity of the stimulus. This mechanism for dy- namically storing information will work for values of parameters such that the system is sensitive to the stimulus, acquiring the desired configuration, yet also able to retain it for some interval of time thereafter. The two main attractors of the system are si = 1 ∀i and si = −1 ∀i. These are the configurations of minimum energy (see the next section for a more detailed discussion on energy). However, the energy is locally minimised for any configuration in which si = dµ(i) ∀i with dµ = ±1; that is, configurations such that each module comprises either all active or all inactive neurons. These are the configurations that we shall use to store information. We define the mean activity4 of each module, mµ ≡ 1 n n Xi∈µ si , 4The mean activity in a neural network model is usually taken to represent the mean firing rate measured in experiments (Torres and Varona, 2010). 86 Chapter 7. Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning which is a mesoscopic variable, as well as the global mean activity, 1 N si = 1 M m ≡ N M Xi Xµ (these magnitudes change with time, but, where possible, we shall avoid writ- ing the time dependence explicitely for clarity). The extent to which the network, at a given time, retains the pattern {ξi} with which it was stimulated is measured with the overlap parameter mµ 1 N 1 M ξisi = mstim ≡ M N Xµ Xi Ideally, the system should be capable of reacting immediately to a stimulus by adopting the right configuration, yet also be able to retain it for long enough to use the information once the stimulus has disappeared. A measure of per- formance for such a task is therefore ξµmµ . 1 τ η ≡ mstim (t), t0+τ Xt=t0+1 where t0 is the time at which the stimulus is received and τ is the period of time we are interested in (η ≤ 1) (Johnson et al., 2008). If the intensity of the stimulus, δ , is very large, then the system will always adopt the right pattern perfectly and η will only depend on how well it can then retain it. In this case, the best network will be one that is made up of unconnected modules. However, since the stimulus in a real brain can be expected to arrive via a relatively small number of axons, either from another part of the brain or directly from sensory cells, it is more realistic to assume that δ is of a similar order as the input a typical neuron receives from its neighbours, hhi ∼ ω hki. Fig. 7.2 shows the mean performance obtained when the network is repeat- edly stimulated with different randomly generated patterns. For low enough values of the modularity λ and stimuli of intensity δ & ω hki, the system can capture and successfully retain any pattern it is “shown” for some period of time, even though this pattern was in no way previously learned. For less in- tense stimuli (δ < ω hki), performance is nonmonotonic with modularity: there exists an optimal value of λ at which the system is sensitive to stimuli yet still able to retain new patterns quite well. It is worth noting that performance can also break down due to thermal fluctuations. The two main attractors of the system (si = 1 ∀i and si = −1 ∀i) 7.3 Cluster Reverberation 87 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 η η η η η η η η η η η η η η η e e e e e e e e e e e e e e e c c c c c c c c c c c c c c c n n n n n n n n n n n n n n n a a a a a a a a a a a a a a a m m m m m m m m m m m m m m m r r r r r r r r r r r r r r r o o o o o o o o o o o o o o o f f f f f f f f f f f f f f f r r r r r r r r r r r r r r r e e e e e e e e e e e e e e e P P P P P P P P P P P P P P P 1 1 1 m m m i i i 0.5 0.5 0.5 t t t s s s m m m 0 0 0 500 500 500 δ=8.5 δ=8.5 δ=8.5 δ=9 δ=9 δ=10 Time (MCS) Time (MCS) Time (MCS) 3500 3500 3500 λ=0.0 λ=0.0 λ=0.0 λ=0.25 λ=0.25 λ=0.5 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Rewiring prob. λ Figure 7.2: Performance η against λ for networks of the sort described in the main text with M = 160 modules of n = 10 neurons, hki = 9; patterns are shown with intensities δ = 8.5, 9 and 10, and T = 0.02 (lines – splines – are drawn as a guide to the eye). Inset: typical time series of mstim (i.e., the overlap with whichever pattern was last shown) for λ = 0.0, 0.25, and 0.5, and δ = hki = 9. suffer the typical second order phase transition of the Hopfield model (Hopfield, 1982), from a memory phase (one in which m = 0 is not stable and stable solutions m 6= 0 exist) to one with no memory (with m = 0 the only stable solution), at the critical temperature (Johnson et al., 2008) Tc = ω hk2 ini hki (Note that, in a directed network, hkini = hkouti ≡ hki, although the other moments can in general be different.) The metastable states we are interested in, though, have a critical temperature . T ′ c = (1 − λ)Tc (assuming that the mean activity of the network is m ≃ 0). That is, the temperature at which the modules are no longer able to retain their individual activity is in general lower than that at which the the solution m = 0 for the whole network becomes stable. 88 Chapter 7. Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning 7.4 Energy and topology Each pair of nodes contributes a configurational energy eij = −ω 1 2 (aij +aj i )sisj ; that is, if there is an edge from i to j and they have opposite activities, the energy is increased in 1 2 ω , whereas it is decreased by the same amount if their activities are the same. Given a configuration, we can obtain its associated energy by summing over all pairs. We shall be interested in configurations with x neurons that have s = +1 (and N − x with s = −1), chosen in such a way that one module at most, say µ, has neurons in both states simultaneously. Therefore, x = nρ + z , where ρ is the number of modules with all their neu- rons in the positive state and z is the number of neurons with positive sign in module µ. We can write m = (2x − 1)/N and mµ = (2z − 1)/n. The total con- figurational energy of the system will be E = Pij eij = 1 2 ω (L↑↓ − hkiN ), where L↑↓ is the number of edges linking nodes with opposite activities. By simply counting over edges, we can obtain the expected value of L↑↓ (which amounts to a mean-field approximation because we are substituting the number of edges between two neurons for its expected value), yielding: + z(n − z) E = (1 − λ) n − 1 ω hki λn 1 N − n {ρ[n − z + n(M − ρ − 1)] + (M − ρ − 1)(z + nρ)} − N . 2 Fig. 7.3 shows the mean-field configurational energy curves for various values of the modularity on a small modular network. The local minima (metastable states) are the configurations used to store patterns. It should be noted that the mapping x → m is highly degenerate: there are C M mM patterns with mean activity m that all have the same energy. (7.1) 7.5 Forgetting avalanches In obtaining the energy we have assumed that the number of synapses rewired from a given module is always ν = hkinλ. However, since each edge is evaluated with probability λ, ν will in fact vary somewhat from one module to another, being approximately Poisson distributed with mean hν i = hkinλ. The depth of the energy well corresponding to a given module is then, neglecting all but the first term in Eq. (7.1) and approximating n − 1 ≃ n, 1 4 ω (nhki − ν ). ∆E ≃ 7.5 Forgetting avalanches 89 M=5 M=5 M=5 M=5 λ=0.2 λ=0.2 λ=0.2 λ=0.2 λ=0.4 λ=0.4 λ=0.4 λ=0.6 λ=0.6 λ=0.8 0 0 0 0 -0.2 -0.2 -0.2 -0.2 -0.4 -0.4 -0.4 -0.4 -0.6 -0.6 -0.6 -0.6 -0.8 -0.8 -0.8 -0.8 ) ) ) ) N N N N k k k k ω ω ω ω ( ( ( ( / / / / E E E E 2 2 2 2 -1 -1 -1 -1 -1 -1 -1 -1 -0.5 -0.5 -0.5 -0.5 0 0 0 0 m m m m 0.5 0.5 0.5 0.5 1 1 1 1 Figure 7.3: Configurational energy of a network composed of M = 20 modules of n = 10 neurons each, according to Eq. (7.1), for various values of the rewiring probability λ. The minima correspond to situations such that all neurons within any given module have the same sign. where (7.2) (7.3) τ −β (τ ) , β (τ ) = 1 + The typical escape time τ from an energy well of depth ∆E at temperature T is τ ∼ e∆E /T (Levine and R.D., 2005). Using Stirling’s approximation in the Poissonian distribution of ν and expressing it in terms of τ , we find that the escape times are distributed according to ln τ (cid:19)− 3 P (τ ) ∼ (cid:18)1 − 4T 2 ωnhki ωnhki "1 + ln λnhki ωnhki ln τ !# . 4T 1 − 4T Therefore, at low temperatures, P (τ ) will behave approximately like a power- law. The left panel of Fig. 7.4 shows the distribution of time intervals between events in which the overlap mµ of at least one module µ changes sign. The power-law-like behaviour is apparent, and justifies talking about forgetting avalanches – since there are cascades of many forgetting events interspersed with long periods of metastability. This is very similar to the behaviour ob- served in other nonequilibrium settings in which power-law statistics arise from the convolution of exponentials (Hurtado et al., 2008; Munoz et al., 2010). It is known from experimental psychology that forgetting in humans is in- deed well described by power-laws (Wixted and Ebbesen, 1991, 1997; Sikstrom, 2002). The right panel of Fig. 7.4 shows the value of the exponent β (τ ) as a function of τ . Although for low temperatures it is almost constant over 90 Chapter 7. Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning ) ) τ τ ( ( p p 100 100 10-1 10-1 10-2 10-2 10-3 10-3 10-4 10-4 101 101 ) ) ) τ τ τ ( ( ( β β β 2 2 2 1.8 1.8 1.8 1.6 1.6 1.6 1.4 1.4 1.4 1.2 1.2 1.2 1 1 1 T=1 T=2 T=3 101 101 101 103 103 103 τ τ τ 105 105 105 107 107 107 102 102 103 103 τ τ 104 104 105 105 Figure 7.4: Left panel: distribution of escape times τ , as defined in the main text, for λ = 0.25 and T = 2. Slope is for β = 1.35. Other parameters as in Fig. 7.2. Symbols from MC simulations and line given by Eqs. (7.2) and (7.3). Right panel: exponent β of the quasi-power-law distribution p(τ ) as given by Eq. (7.3) for temperatures T = 1 (red line), T = 2 (green line) and T = 3 (blue line). many decades of τ – approximating a pure power-law – for any finite T there will always be a τ such that the denominator in the logarithm of Eq. (7.3) approaches zero and β diverges, signifying a truncation of the distribution. 7.6 Clustered networks Although we have illustrated how the mechanism of Cluster Reverberation works on a modular network, it is not actually necessary for the topology to have this characteristic – only for the patterns to be in some way “coarse- grained,” as described, and that each region of the network encoding one bit have a small enough parameter λ, defined as the proportion of synapses to other regions. For instance, for the famous Watts-Strogatz smal l-world model (Watts and Strogatz, 1998) – a ring of N nodes, each initially connected to its k nearest neighbours before a proportion p of the edges are randomly rewired – we have λ ≃ p (which is not surprising considering the resemblance between this model and the modular network used above). More precisely, the expected modularity of a randomly imposed box of n neurons is 2 (cid:19) , n (cid:18) k 1 1 − p 4 − λ = p − n − 1 N − 1 p + 7.6 Clustered networks 91 (7.4) the second term on the right accounting for the edges rewired to the same box, and the third to the edges not rewired but sufficiently close to the border to connect with a different box. Perhaps a more realistic model of clustered network would be a random network embedded in d-dimensional Euclidean space. For this we shall use the scheme laid out by Rozenfeld et al. (Rozenfeld et al., 2002), which consists simply in allocating each node to a site on a d-torus and then, given a par- ticular degree sequence, placing edges to the nearest nodes possible – thereby attempting to minimise total edge length5 . For a scale-free degree sequence (i.e., a set {ki} drawn from a degree distribution p(k) ∼ k−γ ) according to some exponent γ , then, as shown in B, such a network has a modularity 1 d(γ − 2) − 1 (cid:2)d(γ − 2)l−1 − l−d(γ−2) (cid:3) , λ ≃ where l is the linear size of the boxes considered. Fig. 7.5 compares this expression with the value obtained numerically after averaging over many network realizations, and shows that it is fairly good – considering the approximations used for its derivation. It is interesting that even in this scenario, where the boxes of neurons which are to receive the same stimulus are chosen at random with no consideration for the underlying topology, these boxes need not have very many neurons for λ to be quite low (as long as the degree distribution is not too heterogeneous). Carrying out the same repeated stimulation test as on the modular net- works in Fig. 7.2, we find a similar behaviour for the scale-free embedded networks. This is shown in Fig. 7.6, where for high enough intensity of stim- uli δ and scale-free exponent γ , performance can, as in the modular case, be η ≃ 1. We should point out that for good performance on these networks we require more neurons for each bit of information than on modular networks with the same λ (in Fig. 7.6 we use n = 100, as opposed to n = 10 in Fig. 7.2). However, that we should be able to obtain good results for such diverse network topologies underlines that the mechanism of Cluster Reverberation is robust and not dependent on some very specific architecture. In fact, we have recently shown that similar metastable memory states can also occur on networks which have random modularity and clustering, but a certain degree of assortativity 6 (de Franciscis et al., 2011). 5The authors also consider a cutoff distance, but we shall take this to be infinite here. 6The assortativity of a network is here understood to mean the extent to which the degrees of neighbouring nodes are correlated (Johnson et al., 2010b). 92 Chapter 7. Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning λ λ λ λ λ λ λ λ λ 1 1 1 1 1 1 1 1 1 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0 0 0 0 0 0 0 0 0 l=2 l=4 l=8 2 2 2 2 2 2 2 2 2 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 3 3 3 3 3 3 3 3 3 3.5 3.5 3.5 3.5 3.5 3.5 3.5 3.5 3.5 γ γ γ γ γ γ γ γ γ 4 4 4 4 4 4 4 4 4 4.5 4.5 4.5 4.5 4.5 4.5 4.5 4.5 4.5 5 5 5 5 5 5 5 5 5 Figure 7.5: Proportion of outgoing edges, λ, from boxes of linear size l against exponent γ for scale-free networks embedded on 2D lattices. Lines from Eq. (7.4) and symbols from simulations with hki = 4 and N = 1600. δ=3.5 δ=3.5 δ=3.5 δ=3.5 δ=4 δ=4 δ=4 δ=5 δ=5 δ=10 1 1 1 1 1 1 1 1 1 1 1 1 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 η η η η η η η η η η η η e e e e e e e e e e e e c c c c c c c c c c c c n n n n n n n n n n n n a a a a a a a a a a a a m m m m m m m m m m m m r r r r r r r r r r r r o o o o o o o o o o o o f f f f f f f f f f f f r r r r r r r r r r r r e e e e e e e e e e e e P P P P P P P P P P P P γ=2 γ=2 γ=2 γ=3 γ=3 γ=4 1 1 1 0.5 0.5 0.5 m m m i i i t t t s s s m m m 0 0 0 500 500 500 Time (MCS) Time (MCS) Time (MCS) 3000 3000 3000 0 0 0 0 0 0 0 0 0 0 0 0 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4 4 4 4 4 4 SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ SF exponent γ 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 6 6 6 6 6 6 6 Figure 7.6: Performance η against exponent γ for scale-free networks, embed- ded on a 2D lattice, with patterns of M = 16 modules of n = 100 neurons each, hki = 4 and N = 1600; patterns are shown with intensities δ = 3.5, 4, 5 and 10, and T = 0.01 (lines – splines – are drawn as a guide to the eye). Inset: typical time series for γ = 2, 3, and 4, with δ = 5. 7.7 Yes, but does it happen in the brain? 93 7.7 Yes, but does it happen in the brain? As we have shown, Cluster Reverberation is a mechanism available to neural systems for robust short-term memory without synaptic learning. To the best of our knowledge, this is the first mechanism proposed which has these charac- teristics – essential for, say, sensory memory or certain working-memory tasks. All that is needed is for the network topology to be highly clustered or modu- lar, and for small groups of neurons to store one bit of information, as opposed to the conventional view which assumes one bit per neuron. Considering the enormous number of neurons in the brain, and the fact that real individual neurons are probably too noisy to store information reliably, these hypotheses do not seem farfetched. The mechanism is furthermore consistent both with what is known about the topology of the brain, and with experiments which have revealed power-law forgetting. Since the purpose of this paper is only to describe the mechanism of Clus- ter Reverberation, we have made use of the simplest possible model neurons – namely, binary neurons with static, uniform synapses – for the sake of clarity and generality. However, there is no reason to believe that the mechanism would cease to function if more neuronal ingredients were to be incorporated. In fact, cellular bistability, for instance, would increase performance, and the two mechanisms could actually work in conjunction. Similarly, we have also used binary patterns to store information. But it is to be expected that pat- terns depending on any form of frequency coding, for instance, could also be maintained with more sophisticated neurons – such that different modules could be set to different mean frequencies. Whether Cluster Reverberation would work for biological neural systems could be put to the test by growing such modular networks in vitro, stimulating appropriately, and observing the duration of the metastable states. In vivo recordings of neural activity during short-term memory tasks, together with a mapping of the underlying synaptic connections, could be used to ascertain whether the brain does indeed make use of this mechanism – although for this it must be borne in mind that the neurons forming a module need not find themselves close together in metric space. We hope that experiments such as these will be carried out and eventually reveal something more about the basis of this puzzling emergent property of the brain’s known as thought. Chapter 8 Concluding remarks “As long as the brain is a mystery, the universe will remain a mystery,” claimed Santiago Ram´on y Ca jal. Our very essence seems to reside somehow in the workings of this organ, probably as a consequence of electro-chemical signalling that goes on among its hundred billion or so constituent neurons. Will this mystery ever be cleared up? We know of other ob jects that process informa- tion in highly sophisticated ways – electronic computers. Faced with a sudden blue screen, one may be forgiven for calling these devices incomprehensible and capricious, even malevolent. But in fact most educated people understand, on some level at least, what mechanisms and physical processes are behind the complex behaviour displayed by computers, and do not consider the issue a mystery. This is not to suggest that the analogy between brain and computer should be taken any further than to illustrate how a great many elements, each executing some fairly simple and obvious operation, can “cooperate” to yield astonishingly complicated yet functional behaviour; and that one can grasp how this occurs without having to know every detail. But we have not yet reached this point as regards the brain. Much progress has been made con- cerning aspects of physiology, while once unassailable mental disorders such as phobias can now be easily cured by psychology. Yet as far as what mecha- nisms relate these two levels of description goes, perhaps all we can safely say for now is that synaptic plasticity is responsible for long-term memory. The origins of even some well-defined and much studied cognitive abilities – such as probabilistic reasoning or short-term memory – remain somewhat elusive, while the nature of consciousness, say, is still truly a mystery. However, if instead of developing computers ourselves we had been given them by an alien species, we could still hope one day to unravel the mysteries of their magic. In much the same manner, by searching for ways in which collections of neurons 94 95 might perform tasks such as we know them to be capable of, we will some day understand not only how our stomachs digest and our hearts pump, but also how our brains think. I cannot pretend that the work described here takes us more than, at best, a tiny step of the way along this path. The brain is, among other things, a network, and networks are a kind of mathematical ob ject about which we now know much more than just a few years ago. In fact, they are a central element of what can arguably be called the most challenging frontier currently facing human understanding about the world – the nature of complex systems. So, from among the innumerable aspects likely to shape and determine the way neurons cooperate, the research presented here focuses on the structure of the underlying network. First of all it looks at how this structure can develop. Chapter 3 addressed this by formalizing as a stochastic process a situation governed by probabilistic events like synaptic growth and death. Such simple individual behaviour was shown to be enough to explain many statistical features of real neural systems. Furthermore, this Fokker-Planck description relating microscopic, stochastic actions to a macroscopic evolution of properties such as mean synaptic density, degree heterogeneity or assortativity may help to gain insights into the biochemical processes taking place. The rest of the thesis is mostly devoted to how aspects of a neural network’s topology might influence or even determine its ability to carry out certain tasks akin to those the brain undertakes. The fact that dynamical memory perfor- mance ensuing from synaptic depression is favoured by a highly heterogeneous degree distribution, laid out in Chapter 4, may help to explain why the brain seems to display such a topology at several levels of description – perhaps somehow maintaining itself close to a critical point. Similarly, the enhanced robustness to noise found for positively correlated networks in Chapter 6 sug- gests a functional advantage to a neural network being thus wired; a prediction also in agreement with some experimental findings. As far as unearthing the mechanisms underpinning how neurons can per- form cognitive tasks goes, though, perhaps the most interesting idea proposed is that of Cluster Reverberation, in Chapter 7, whereby thanks to modularity and/or clustering a neural network is able to store information instantly, with- out requiring biochemical changes in the synapses. Time will tell whether real neural systems do indeed harness this mechanism to perform certain short-term memory tasks. A collateral but noteworthy aspect of this research is the potentiality for 96 Chapter 8. Concluding remarks application elsewhere of some of the mathematical techniques developed. Most of all, the method for studying correlated networks and dynamics thereon put forward in Chapter 5 for use in Chapter 6 can be expected to find widespread use. The answer to the question of why most networks are disassortative given in Chapter 5, or the relation between degree-degree correlations and nestedness described in Appendix C are examples of this. Finally I must mention not just the answers I hope to have provided, or at least hinted at, to some unsolved problems, but also the questions that have been posed and challenges laid bare: Would a more detailed description of brain development still be possible with Fokker-Planck equations? Are these topological effects, found to be at work for the simplest neural models, indeed so relevant for real neurons? Can Cluster Reverberation be performed in vitro? The greatest function this thesis could perform would be to stimulate others to look into these or related issues in more depth than here. But I also hope it may serve to illustrate the sentiment, What matters how long the path to the final unravelling of the mysteries is, as long as the going is fun? Chapter 9 Conclusiones en espanol “Mientras el cerebro sea un misterio, el universo continuar´a siendo un miste- rio”, dijo una vez Santiago Ram´on y Ca jal. Parece que nuestra misma esencia reside de alguna manera en el funcionamiento de este ´organo, probablemente como consecuencia de las senales electro-qu´ımicas entre sus aproximadamente cien mil millones de neuronas. ¿Se resolver´a alg´un d´ıa este misterio? Conoce- mos otros ob jetos capaces de procesar informaci´on de manera altamente sofisti- cada: los ordenadores electr´onicos. Confrontados con un pantallazo azul, se nos podr´ıa perdonar el tildar este tipo de aparatos de incomprensibles y capri- chosos, por no decir mal´evolos. Pero en realidad la mayor parte de la gente entiende, al menos en alg´un nivel, cu´ales son los mecanismos y procesos f´ısicos que subyacen el comportamiento complejo del que hacen gala los ordenadores, y no consideran que el tema sea un misterio. No es que la analog´ıa entre cerebro y ordenador deba ser llevado m´as lejos que para ilustrar c´omo muchos elementos, cada uno ejecutando alguna operaci´on relativamente simple y obvia, pueden “cooperar” y mostrar un comportamiento colectivo asombrosamente complicado, pero funcional; y que se puede comprender c´omo ocurre esto sin necesidad de conocer hasta el ´ultimo detalle. A´un no hemos llegado a poder responder a esta pregunta en lo que respecta al cerebro. Hemos ampliado enormemente nuestro conocimiento de aspectos fisiol´ogicos, y trastornos men- tales antano incurables, como las fobias, son f´acilmente tratadas hoy en d´ıa por la psicolog´ıa. En cuanto a los mecanismos que relacionan estos dos niveles de descripci´on, posiblemente lo ´unico que podamos decir a ciencia cierta es que la plasticidad sin´aptica est´a detr´as de la memoria a largo plazo. Los or´ıgenes incluso de algunas habilidades cognitivas bien definidas y extensamente estu- diadas, como el razonamiento probabil´ıstico o la memoria a corto plazo, est´an a´un por descifrar completamente; mientras que, por ejemplo, la naturaleza de 98 99 la consciencia es verdaderamente a´un un misterio. Sin embargo, si en lugar de haber desarrollado los ordenadores nosotros mismos los hubi´esemos recibido de una especie alien´ıgena, a´un as´ı podr´ıamos esperar alg´un d´ıa desenmaranar los misterios de su magia. Del mismo modo, buscando maneras de que conjuntos de neuronas puedan realizar el tipo de tareas de las que las sabemos capaces, alg´un d´ıa entenderemos no s´olo c´omo nuestros est´omagos digieren y nuestros corazones laten, sino tambi´en c´omo nuestros cerebros piensan. Este traba jo, en el mejor de los casos, nos avanza un paso infinitesimal por este camino. El cerebro es, entre otras muchas cosas, una red, y las redes son ob jetos matem´aticos sobre los que sabemos hoy mucho m´as que hace tan s´olo unos anos. De hecho, son un elemento fundamental para uno de los mayores re- tos con los que se enfrenta actualmente el conocimiento humano: la naturaleza de los sistemas complejos. As´ı que, de entre los innumerables aspectos suscepti- bles de modificar y determinar c´omo las neuronas cooperan, esta investigaci´on se centra en la estructura de la red subyacente. Primero analiza c´omo dicha estructura puede desarrollarse. El Cap´ıtulo 3 enfoca esto formalizando medi- ante la teor´ıa de los procesos estoc´asticos una situaci´on gobernada por eventos probabil´ısticos tales como el crecimiento y la muerte sin´apticas. Se demuestra que este tipo de comportamiento individual es suficiente para explicar muchas propiedades estad´ısticas de las redes de cerebros reales. Por otra parte, este marco te´orico puede ser reducido a una descripci´on en t´erminos de ecuaciones de Fokker-Planck, que relacionan acciones microsc´opicas estoc´asticas con la evoluci´on macrosc´pica de propiedades como la densidad sin´aptica media, la heterogeneidad de la distribuci´on de grados o la asortatividad, que quiz´as nos permita extraer informaci´on relevante acerca de los procesos bioqu´ımicos in- volucrados. La mayor parte del resto de la tesis trata de c´omo aspectos de la topolog´ıa de una red neuronal pueden influenciar o incluso determinar su habilidad para ejecutar ciertas tareas cognitivas como las que se describen en un cerebro o medio neuronal real. Por ejemplo, el hecho de que, en cuanto a la memoria din´amica que emerge gracias a la depresi´on sin´aptica, el rendimiento es mayor para una distribuci´on de grados altamente heterog´enea, como demuestra el Cap´ıtulo 4, podr´ıa ayudar a explicar por qu´e el cerebro parece mostrar una topolog´ıa de este tipo en varios niveles de descripci´on, quiz´as incluso man- teni´endo su actividad, de alguna manera todav´ıa no comprendida del todo, cerca de un punto cr´ıtico. De igual modo, la mayor robustez durante los pro- cesos cognitivos en presencia de ruido en el caso de redes con correlaciones 100 Chapter 9. Conclusiones en espanol positivas como se ha descrito en el Cap´ıtulo 6 sugiere que existe una venta ja funcional para una red neuronal en adoptar esta propiedad; una predicci´on que tambi´en enca ja con algunos hallazgos experimentales. En lo que se refiere a desentranar los mecanismos que permiten a las neuronas realizar colectivamente tareas cognitivas, quiz´as la idea m´as intere- sante aqu´ı propuesta es la de Cluster Reverberation (Reverberaci´on de Grupo), en el Cap´ıtulo 7, seg´un la cual, gracias a la modularidad y/o el grado de “agrupamiento”, una red neuronal es capaz de almacenar informaci´on in- stant´aneamente, sin requerir para ello cambios bioqu´ımicos de potenciaci´on o depresi´on a largo plazo en las sinapsis. El tiempo dir´a si el cerebro aprovecha realmente este mecanismo para realizar ciertas tareas de memoria de corto plazo. Un aspecto colateral pero digno de menci´on de este traba jo es el de la potencialidad de algunas de las t´ecnicas matem´aticas desarrolladas de ser apli- cadas para otras situaciones de inter´es. Sobre todo, es de esperar que el m´etodo para estudiar redes correlacionadas, y din´amicas sobre ellas, propuesto en el Cap´ıtulo 5 y utilizado en el Cap´ıtulo 6, sea ´util para una amplia gama de prob- lemas. La respuesta, en el Cap´ıtulo 5, a la pregunta de por qu´e la mayor´ıa de las redes son disasortativas, o la relaci´on entre correlaciones entre los nodos y el “anidamiento” descrita en el Ap´endice C son ejemplos de aplicaciones. Finalmente, hay que mencionar no s´olo las respuestas que se han intentado dar, o al menos sugerir, con esta tesis para algunos problemas sin resolver, sino tambi´en las preguntas y los nuevos retos que han surgido: por ejemplo, ¿ser´ıa posible, tambi´en con ecuaciones de Fokker-Planck, una descripci´on m´as detallada del desarrollo cerebral? ¿Son estos efectos topol´ogicos, descritos para los modelos neuronales m´as sencillos, realmente tan relevantes para neuronas de verdad? ¿Puede el mecanismo de Cluster Reverberation ocurrir in vitro? En definitiva, la mayor funci´on que pudiera cumplir esta tesis ser´ıa la de estimular a otra/os para que indaguen en estos y otros temas m´as profundamente que aqu´ı. Pero quiz´as tambi´en sirva para ilustrar el siguiente sentimiento: ¿qu´e m´as da cu´an largo sea el camino hacia el desenmaranamiento ´ultimo de los misterios, siempre que el trayecto sea divertido? Appendix A Nonlinear preferential rewiring in fixed-size networks as a diffusion process We present an evolving network model in which the total numbers of nodes and edges are conserved, but in which edges are continuously rewired according to nonlinear preferential detachment and reattachment. Assuming power-law kernels with exponents α and β , the stationary states the degree distribu- tions evolve towards exhibit a second order phase transition – from relatively homogeneous to highly heterogeneous (with the emergence of starlike struc- tures) at α = β . Temporal evolution of the distribution in this critical regime is shown to follow a nonlinear diffusion equation, arriving at either pure or mixed power-laws, of exponents −α and 1 − α. Complex systems may often be described as a set of nodes with edges con- necting some of them – the neighbours – (see, for instance, Refs.(Boccaletti et al., 2006; Arenas et al., 2008a; Marro et al., 2008)). The number of edges a partic- ular node has is called its degree, k . The study of such large networks is usually made simpler by considering statistical properties, e.g., the degree distribution, p(k) (probability of finding a node with a particular degree). It turns out that a high proportion of real-world networks follow power-law degree distributions, p(k) ∼ k−γ – referred to as scale-free due to their lack of a characteristic size. Also, many of them have their edges placed among the nodes apparently in a random way – i.e., there is no correlation between the degree of a node and any other of its properties, such as the degrees of its neighbours. Barab´asi and Albert (Barab´asi and Albert, 1999) applied the mechanism of preferential attachment to an evolving network model and showed how this resulted in the 102 103 degree distributions becoming scale-free for long enough times. For this to work, attachment had to be linear – i.e., the probability a node with degree k has of receiving a new edge is π(k) ∼ k + q . This results in scale-free stationary degree distributions with an exponent γ = 3 − q . Preferential attachment seems to be behind the emergence of many real- world, continuously growing networks. However, not all networks in which some nodes at times gain (or loose) new edges have a continuously growing number of nodes. For example, a given group of people may form an evolv- ing social network (Kossinets and Watts, 2006) in which the edges represent friendship. Preferential attachment may be relevant here – the more people you know, the more likely it is that you will be introduced to someone new – but probabilities are not expected to depend linearly on degree. For instance, there may be saturations (highly connected people might become less acces- sible), threshold effects (hermits may be prone to antisocial tendencies), and other non-linearities. The brain may also be a relevant case. Once formed, the number of neurons does not seem to continually augment, and yet its struc- tural topology is dynamic (Klintsova and Greenough, 1999). Synaptic growth and dendritic arborization have been shown to increase with electric stimula- tion (Lee et al., 1980; Roo et al., 2008) – and, in general, the more connected a neuron is, the more current it receives from the sum of its neighbours. Barab´asi and Albert showed that both (linear) preferential attachment and an ever-growing number of nodes are needed for scaling to emerge in their model. In a fixed population, their mechanism would result in a fully-connected network. However, this is not normally observed in real systems. Rather, just as some new edges sprout, others disappear – less used synapses suffer at- rophy, unstimulating friendships wither. Often, the numbers of both nodes and edges remain roughly constant. The same authors did therefore extend their model so as to include the effects of preferential rewiring (which could be applied to fixed-size networks), although again probabilities depended lin- early on node degree (Albert and Barab´asi, 2000). Another mechanism which (roughly) maintains constant the numbers of nodes and edges is node fus- ing (Thurner et al., 2007), once more according to linear probabilities. As to nonlinear preferential attachment, the (growing) BA model was extended to take power-law probabilities into account (Krapivsky et al., 2000), although the solutions are only scale free for the linear case. In this note we present an evolving network model with preferential rewiring according to nonlinear (power-law) probabilities. The number of nodes and Chapter A. Nonlinear preferential rewiring in fixed-size networks 104 as a diffusion process edges is conserved but the topology evolves, arriving eventually at a macro- scopically (nonequilibrium) stationary state – as described by global properties such as the degree distribution. Depending on the exponents chosen for the rewiring probabilities, the final state can be either fairly homogeneous, with a typical size, or highly heterogeneous, with the emergence of starlike struc- tures. In the critical case marking the transition between these two regimes, the degree distribution is shown to follow a nonlinear diffusion equation. This describes a tendency towards stationary states that are characterized either by scale-free or by mixed scale-free distributions, depending on parameters. Our model consists of a random network with N nodes of respective de- gree ki , i = 1, 2, ..., N , and 1 2 N hki edges. Initially, the degrees have a given distribution p(k , t = 0). At each time step, one node is chosen with a proba- bility which is a function of its degree, ρ(ki ). One of its edges is then chosen randomly and removed from it, to be reconnected to another node j chosen according to a probability π(kj ). That is, an edge is broken and another one is created, and the total number of edges, as well as the total number of nodes, is conserved. The functions π(k) and ρ(k) are arbitrary, but we shall explic- i and ρ(ki ) ∼ kβ itly illustrate here π(ki ) ∼ kα i that capture the essence of a wide class of nonlinear monotonous response functions and are easy to handle analytically. The probabilities π and ρ a given node has, at each time step, of increasing or decreasing its degree can be interpreted as transition probabilities between states. The expected value of the increment in a given p(k , t) at each time step, ∆p(k , t), may then be written as ∂p(k , t) ∂ t (A.1) = (k − 1)α ¯k−1 α p(k − 1, t) + (k + 1)β ¯k−1 β p(k + 1, t) − (cid:0)kα ¯k−1 α + kβ ¯k−1 β (cid:1) p(k , t), where ¯ka = ¯ka (t) = Pk kap(k , t). If it exists, any stationary solution must satisfy the condition pst(k + 1) (k + 1)β ¯k st α = pst (k) kα ¯k st β which, for k ≫ 1, implies that = ¯k st (k + 1)β − 1! pst(k). kα ∂pst (k) α ¯k st ∂k β 1 Therefore, the distribution will have an extremum at ke = (cid:0)¯k st β /¯k st α (cid:1) α−β (where If α < β , this will be a maximum, we have approximated ke ≃ ke + 1). (A.2) 105 signalling the peak of the distribution. On the other hand, if α > β , ke will correspond to a minimum. Therefore, most of the distribution will be broken in two parts, one for k < ke and another for k > ke . The critical case for α = β will correspond to a monotonously decreasing stationary distribution, but such that limk→∞∂pst (k)/∂k = 0. In fact, Eq. (A.1) is for this situation (α = β ) the discretized version of a nonlinear diffusion equation, ∂ 2 ∂k2 [kαp(k , τ )], after dynamically modifying the time scale according to τ = t/¯kα (t). Ignoring, for the moment, border effects, the solutions of this equation are of the form ∂p(k , τ ) ∂ τ = (A.3) (A.4) pst (k) ∼ Ak−α + Bk−α+1 , with A and B constants. If α > 2, then given A we can always find a B which allows pst (k) to be normalized in the thermodynamic limit 1 . For example, if the lower limit is k ≥ 1, then B = (α − 2) [1 − A/(α − 1)]. However, if 1 < α ≤ 2, then only A can remain non-zero, and pst (k) will be a pure power law. For α ≤ 1, both constants must tend to zero as N → ∞. In finite networks, no node can have a degree larger than N − 1 or lower than 0. In fact, one would usually wish to impose a minimum nonzero degree, e.g. k ≥ 1. The temporal evolution of the degree distribution is illustrated in Fig. A.1. This shows the result of integrating Eq. (A.1) for k ≥ 1, different times, β = 1, and three different values of α, along with the respective values obtained from Monte Carlo simulations. The main result may be summarized as follows. For α < β , the network will evolve to have a characteristic size, centred around hki. At the critical case α = β , all sizes appear, according either to a pure or a composite power law, as detailed above. If we impose, say, k ≥ 1, then starlike structures will emerge, with a great many nodes connected to just a few hubs 2 . Figure A.2 illustrates the second order phase transition undergone by the variance of the final (stationary) degree distribution, depending on the expo- nent α, where β is set to unity. It should be mentioned that this particular 1Although all moments of k will diverge unless B = 0. 2There is a finite-size effect not taken into account by the theory – but relevant when α > β – which provides a natural lower cutoff for pst (k): if there are, say, m nodes which are connected to the whole network, then the minimum degree a node can have is m. Chapter A. Nonlinear preferential rewiring in fixed-size networks 106 as a diffusion process 100 100 100 100 100 100 100 100 α=0.5 α=0.5 α=0.5 α=0.5 α=0.5 α=0.5 α=0.5 α=0.5 ) ) ) ) ) ) ) ) k k k k k k k k ( ( ( ( ( ( ( ( t t t t t t t t s s s s s s s s p p p p p p p p ) ) ) ) ) ) ) ) ) k k k k k k k k k ( ( ( ( ( ( ( ( ( t t t t t t t t t s s s s s s s s s p p p p p p p p p ) ) ) ) ) ) ) ) k k k k k k k k ( ( ( ( ( ( ( ( t t t t t t t t s s s s s s s s p p p p p p p p 100 100 100 100 100 100 100 100 104 104 104 104 104 104 104 104 100 100 100 100 100 100 100 100 100 104 104 104 104 104 104 104 104 104 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 104 104 104 104 104 104 104 104 100 100 100 100 100 100 100 100 t=102 t=102 t=102 t=102 t=103 t=103 t=103 t=104 t=104 t=105 k k k k k k k k 102 102 102 102 102 102 102 102 α=1.0 α=1.0 α=1.0 α=1.0 α=1.0 α=1.0 α=1.0 α=1.0 α=1.0 1/k α=1.5 α=1.5 α=1.5 α=1.5 α=1.5 α=1.5 α=1.5 α=1.5 k k k k k k k k k k k k k k k k k 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 Figure A.1: Degree distribution p(k , t) at four different stages of evolution: t = 102 [(yellow) squares], 103 [(blue) circles], 104 [(red) triangles)] and 105 MCS [(black) diamonds]. From top to bottom panels, subcritical (α = 0.5), critical (α = 1) and supercritical (α = 1.5) rewiring exponents. Symbols from MC simulations and corresponding solid lines from numerical integration of Eq. (A.1). β = 1, hki = 10 and N = 1000 in all cases. case, β = 1, corresponds to edges being chosen at random for disconnection, since the probability of a random edge belonging to node i is proportional to ki . This topological phase transition is similar to the ones that have been 107 100 100 100 100 10 10 10 10 10 10 10 10 10 10 10 10 ) k ( t s p 2 2 2 2 2 2 2 2 2 2 2 2 > > > > > > > > > > > > k k k k k k k k k k k k < < < < < < < < < < < < / / / / / / / / / / / / 2 2 2 2 2 2 2 2 2 2 2 2 σ σ σ σ σ σ σ σ σ σ σ σ 10-3 10-3 10-3 10-3 100 100 100 100 N=800 N=1200 N=1600 N=2000 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 α=0.5 α=0.5 α=0.5 α=0.5 α=1.0 α=1.0 α=1.0 α=1.5 α=1.5 β=1 β=1 β=1 β=1 β=1 β=1 β=1 β=1 β=1 β=1 β=1 β=1 103 103 103 103 k 10 10 10 1 1 1 2 2 2 > > > k k k < < < / / / 2 2 2 σ σ σ 0.1 0.1 0.1 1 1 1 1 1 1 1 1 1 1 1 1 α α α α α α α α α α α α 105 105 105 0 0 0 Time (in MCS) Time (in MCS) Time (in MCS) 2 2 2 2 2 2 2 2 2 2 2 2 Figure A.2: Adjusted variance σ 2/hki2 of the degree distribution after 2 × 105 MCS against α, as obtained from MC simulations, for system sizes N = 800 [(yellow) squares], 1200 [(blue) circles], 1600 [(red) triangles] and 2000 [(black) diamonds]. Top left inset shows final degree distributions for α = 0.5 [light gray (blue)], 1 [dark gray (red)] and 1.5 (black), with N = 1000. Bottom right inset shows typical time series of σ 2/hki2 for the same three values of α and N = 1200. In all cases, β = 1 and hki = 10. described in equilibrium network ensembles defined via an energy function, in the so-called synchronic approach to network analysis (Farkas et al., 2004; Park and Newman, 2004; Burda et al., 2004; Der´enyi et al., 2004). However, our (nonequilibrium) model does not come within the scope of this body of work, since the rewiring rates cannot, in general, be derived from a potential. Furthermore, we are here concerned with the time evolution rather than the stationary states, making our approach diachronic. Summing up, in spite of its simplicity, our model captures the essence of many real-world networks which evolve while leaving the total numbers of nodes and edges roughly constant. The grade of heterogeneity of the station- ary distribution obtained is seen to depend crucially on the relation between the exponents modelling the probabilities a node has of obtaining or loos- ing a new edge. It is worth mentioning that the heterogeneity of the degree distribution of a random network has been found to determine many rele- vant behaviours and magnitudes such as its clustering coefficient and mean minimum path (Newman, 2003c), critical values related to the dynamics of Chapter A. Nonlinear preferential rewiring in fixed-size networks 108 as a diffusion process excitable networks (Johnson et al., 2008), or the synchronizability for systems of coupled oscillators (since this depends on the spectral gap of the Laplacian matrix) (Barahona and Pecora, 2002). The above shows how scale-free distributions, with a range of exponents, may emerge for nonlinear rewiring, although only in the critical situation in which the probabilities of gaining or loosing edges are the same. We believe that this non-trivial relation between the microscopic rewiring actions (gov- erned in our case by parameters α and β ) and the emergent macroscopic degree distributions could shed light on a class of biological, social and communica- tions networks. Appendix B Effective modularity of highly clustered networks The number of nodes within a radius r is n(r) = Adrd , with Ad a constant. We shall therefore assume a node with degree k to have edges to all nodes up to a distance r(k) = (k/Ad)1/d , and none beyond (note that this is not necessarily always feasible in practice). To estimate λ, we shall first calcu- late the probability that a randomly chosen edge have length x. The chance that the edge belong to a node with degree k is π(k) ∼ kp(k) (where p(k) is the degree distribution). The proportion of edges that have length x among those belonging to a node with degree k is ν (xk) = dAdxd−1/k if Adxd < k , and 0 otherwise. Considering, for example, scale-free networks (as in Ref. (Rozenfeld et al., 2002)), so that the degree distribution is p(k) ∼ k−γ in some interval k ∈ [k0 , kmax ] (Barab´asi and Albert, 1999), and integrating over p(k), we have the distribution of lengths, P (x) = (C onst.) Z kmax max(k0 ,Axd ) where we have assumed, for simplicity, that the network is sufficiently sparse that max(k0 , Axd ) = Axd , ∀x ≥ 1, and where we have normalised for the interval 1 ≤ x < ∞; strictly, x ≤ (kmax/A)1/d , but we shall also ignore this effect. Next we need the probability that an edge of length x fall between two compartments of linear size l. This depends on the geometry of the situation as well as dimensionality; however, a first approximation which is independent of such considerations is π(k)ν (k x)dk = d(γ − 2)x−[d(γ−2)+1] , Pout (x) = min (cid:16)1, 110 x l (cid:17) . 111 We can now estimate the modularity λ as λ = Z ∞ 1 d(γ − 2) − 1 (cid:2)d(γ − 2)l−1 − l−d(γ−2) (cid:3) . 1 Fig. 7.5 shows how λ depends on γ for d = 2 and various box sizes. Pout (x)P (x)dx = Appendix C Nestedness of networks The property of nestedness has for some time aroused a fair amount of interest as regards ecological networks – especially since a high nestedness in mutu- alistic systems has been shown to enhance biodiversity. However, because it is usually estimated with software, no analytical work has been done relating nestedness with other network characteristics, and consequently comparisons of experimental data with null-models can only be done computationally. We suggest a slightly refined version of the measure recently defined by Bastolla et al. and go on to study the effect of the degree distribution and degree correla- tions (assortativity). Our work provides a benchmark against which empirical networks can be contrasted. C.1 Introduction The intense study that complex networks have undergone over the past decade or so has shown how important topological features can be for properties of complex systems, such as dynamical behaviour, spreading of information, re- silience to attacks, etc. (Newman, 2003c; Boccaletti et al., 2006). A paradig- matic case is that of ecosystems. The solution to May’s paradox (May, 1973) – the fact that large ecosystems seem to be especially stable, when theory predicts the contrary – is still not clear, but it is widely suspected that there is some structural feature of ecological networks which as yet eludes us. One aspect of such networks, which has been studied for some time by ecologists and may be related to this problem, is called nestedness. Loosly speaking, a network – say of species and islands, linked whenever the former inhabit the latter – is said to be highly nested if the species which exist on scarcely popu- lated islands tend always to be found also on those islands inhabited by many 112 C.1 Introduction 113 Figure C.1: Maximally packed matrix representing a network of plants and islands off Perth (Abbott and Black, 1980) (because the network is bipartite, the adjacency matrix is composed of four blocks: two identical to this ma- trix, the other two composed of zeros). Data, image and line obtained from NESTEDNESS CALCULATOR, which returns a “temperature” of T = 0.69o for this particular network. different species. This can be most easily seen by graphically representing a matrix such that animals are columns and islands are files, with elements equal to one whenever two nodes are linked and zero if not. If, after ordering each kind of node by degree (number of neighbours), all the ones can be quite neatly packed into one corner, the network is considered highly nested. This is done in Fig. C.1 for a network of plants inhabiting islands off Perth. This rather vague concept is usually measured with software for the purpose. For Fig. C.1, we have used NESTEDNESS CALCULATOR, which estimates a curve of equal density of ones and zeros, calculates how many ones and zeros are on the “wrong” side and by how much, and returns a number between 0 and 100 called “temperature” by analogy with some system such as a subliming solid. A low temperature indicates high nestedness. To determine how significantly nested a given network is, the usual procedure is to generate equivalent ran- dom networks computationally (with sone constraint such as the number of edges or the degree of each node being conserved) and estimate how likely it is that such a network be “colder” than that of the data. Bastolla et al. (Bastolla et al., 2009) have recently shown how symbiotic interactions can reduce the effective competition between two species, say of insect, via common symbiotic hosts – such as plants they pollinate. These authors define a measure to take into account the average number of shared partners in these mutualistic networks, and call it “nestedness” because it would seem to be related to the concept referred to above. They go on to show 114 Chapter C. Nestedness of networks evidence of how the nestedness of empirical mutualistic networks is correlated with the biodiversity of the corresponding ecosystems. This beneficial effect “enemy” nodes can gain from sharing “friendly” partners is not confined to ecosystems. It is expected also to play a role, for instance, in financial networks or other economic systems (Sugihara and Ye, 2009). The principle is simple. Say nodes A and B are in competition with each other. An increase in A will be to B’s detriment, and viceversa; but if both A and B engage in a symbiotic relationship with node C, then A’s thriving will stimulate C, which in turn will be helpful to B. Thus, the effective competition between A and B is reduced, and the whole system becomes more stable and capable of sustaining more nodes (Dom´ınguez-Chibet´ın et al., 2011). In Ref. (Johnson and Munoz) we take up this idea of shared neighbours (though characterised with a slightly different measure, for reasons we shall explain in Section C.2) and study analytically the effect of other topological properties, such as the degree distribution and degree-degree correlations. This allows us to contrast empirical data with null-models and thus test for statis- tical significance with no need of computer randomisations. We also comment on how mutual-neighbour structure could develop in systems of interdependent networks (such as competition and symbiosis) so as to minimise the risk of a “cascade of failures” (Buldyrev et al., 2010). Although we are not here con- cerned specifically with neural systems, a description of this work is included as an appendix since it serves as an example application of the method put forward in Ref. (Johnson et al., 2010b) and presented in Chapter 5. C.2 Definition Consider a network with N nodes defined by the adjacency matrix a: the element aij is equal to the number on links, or edges, from node j to node i (typically considered to be either one or zero). If a is symmetric, then the network is undirected and each node i can be characterised by a degree ki = Pj aij . (If it is directed, i has both an in degree, k in i = Pj aij , and an out degree, kout = Pj aj i ; we shall focus here on undirected networks, although i most of the results could be easily extended to directed ones.). Bastolla et al. (Bastolla et al., 2009) have shown that the effective compe- tition between two species (say two species of insect) can be reduced if they have common neighbours with which they are in symbiosis (for instance, if they both pollinate the same plant). Therefore, in mutualistic networks (net- C.3 The effect of the degree distribution 115 , (C.1) works of symbiotic interactions) it is beneficial to the species at two nodes i and j for the number of shared symbiotic partners, nij = Pl ail alj = (a2 )ij , to be high. Going on this, and assuming the network is undirected, the authors suggest taking into account the following measure: ηB = Pi<j nij Pi<j min(ki , kj ) which they call nestedness because it would seem to be highly correlated with the measures returned by nestedness software. Note that, although the authors were considering only bipartite graphs, this characteristic is not imposed in the above definition. In this work, we shall take up the idea of the importance of nij , but use a slightly different measure of nestedness, for several reasons. One is that ηB has a serious shortcoming. If we commute the sums1 in the numer- ator of Eq. (C.1), we find that the result only depends on the heterogeneity of the degree distribution: Pij nij = Pl Pi ail Pj alj = N hk2 i. Also, although the maximum value nij can take is min(ki , kj ), this is not necessarily the best normalisation factor, since the expected number of paths of length 2 connect- ing nodes i and j depends on both ki and kj (as we show explicitely in Section C.3). Furthermore, it can sometimes be convenient to have a local measure of nestedness. For these reasons, we shall use ηij ≡ nij kikj = (a2)ij kikj , (C.2) which is defined for every pair of nodes (i, j ). This allows for the consideration of a nestedness per node, ηi = N −1 Pj ηij , or of the global measure 1 N 2 Xij ηij . η = C.3 The effect of the degree distribution (C.3) Most networks have quite broad degree distributions p(k), most notably the fairly ubiquitous scale-free networks, for which they follow power-laws, p(k) ∼ k−γ . Since this heterogeneity tends to have an importante influence on any network measure, it will be useful to take this effect into account analytically. As is standard, the null-model we shall use to do this is the configurational 1 In an undirected network, Pi<j = 1 2 Pij ; we shall always sum over all i and j , since it is easier to generalise to directed networks and often avoids writing factors 2. 116 Chapter C. Nestedness of networks model (Newman, 2003c): the set of random networks wired according to the constraints that a given degree sequence (k1 , ..., kN ) is respected, and also that there be no degree-degree correlations. The expected value of an element of the adjacency matrix for networks belonging to this ensemble is . aij ≡ ǫc ij = kikj hkiN We shall use a line, (·), to represent expected values given certain constraints, and angles, h·i, for averages over nodes of a given network2 . For the case of the adjacency matrix, we use the notation ǫc ij = aij for clarity and coherence with previous work. Plugging Eq. (C.4) into Eq. (C.2), we have the expected value in the configuration ensemble, (C.4) (C.5) ηij = hk2i hki2N ≡ ηconf . Since ηc is independent of i and j , it coincides with the expected value for the global measure, η = ηconf – a fact that justifies the normalisation chosen in Eq. (C.2). It is obvious from Eq. (C.5) that degree heterogeneity will have an important effect on η . Therefore, if we are to capture aspects of network structure other than those directly induced by the degree distribution, it will in general be useful to consider the nestedness normalised to this expected value, η ≡ η ηconf = hki2 hk2iN Xij Although η is unbounded, it has the advantage that it is equal to unity for any uncorrelated random network, independently of its degree heterogeneity, thereby making it possible to detect non-trivial structure in a given empirical network without the need for computational randomisations. (a2)ij kikj . (C.6) C.4 Nestedness and assortativity In the configuration ensemble, the expected value of the mean degree of the neighbours of a given node is knn,i = k−1 i Pj ǫc ij kj = hk2i/hki, which is indepen- dent of ki . However, real networks usually display degree-degree correlations, If knn (k) increases (decreases) with k , with the result that knn,i = knn (ki ). 2 In this case, for instance, the network considered for hki is any of the members of the ensemble, since they all have the same mean degree by definition. C.4 Nestedness and assortativity 117 the network is assortative (disassortative). A measure of this phenomenon is Pearson’s coefficient applied to the edges (Newman, 2003c, 2002, 2003a; l ] − [kl ]2 ), where kl and k ′ l ] − [kl ]2 )/([k2 Boccaletti et al., 2006): r = ([klk ′ l are the degrees of each of the two nodes belonging to edge l, and [·] ≡ (hkiN )−1 Pl (·) is an average over edges. Writing Pl (·) = Pij aij (·), r can be expressed as r = hkihk2knn(k)i − hk2i2 . (C.7) hkihk3i − hk2i2 The ensemble of all networks with a given degree sequence (k1 , ...kN ) con- tains a subset for all members of which knn (k) is constant (the configuration ensemble), but also subsets displaying other functions knn (k). In Chapter 5 (Johnson et al., 2010b) we showed that there is a one-to-one mapping between any mean-nearest-neighbour function knn(k) and its corre- sponding mean-adjacency-matrix ǫ, which is as follows: writing knn (k) as + Z dν f (ν )σν+1 (cid:20) kν−1 knn(k) = hk2 i k (cid:21) 1 hkν i − hki with σν+1 ≡ hkν+1i − hkihkν i (which can always be done), the corresponding matrix ǫ takes the form (C.8) ǫij = (C.9) N (cid:20) (kikj )ν j + hkν i(cid:21) . kikj f (ν ) + Z dν hkν i − kν i − kν hkiN In many empirical networks, knn (k) has the form knn (k) = A + Bkβ , with A, B > 0 (Boccaletti et al., 2006; Pastor-Satorras et al., 2001) – the mixing being assortative (disassortative) if β is positive (negative). Such a case is fitted by Eq. (C.8) if f (ν ) = C [δ (ν − β − 1)σ2/σβ+2 − δ (ν − 1)], with C a positive constant, since this choice yields + C σ2 (cid:20) kβ knn (k) = hk2i hki (cid:21) . 1 hkβ+1i − hki After plugging Eq. (C.10) into Eq. (C.7), one obtains: hkβ+1i (cid:18) hkihkβ+2i − hk2 ihkβ+1i (cid:19) . C σ2 hkihk3i − hk2i2 It turns out that the configurations most likely to arise naturally (those with maximum entropy) usually have C ≃ 1 (Johnson et al., 2010b) (c.f. Chapter 5). Therefore, and for the sake of analytical tractability, we shall do as in (C.11) (C.10) r = 118 Chapter C. Nestedness of networks ǫij = Chapter 6 and consider this particular case3 – that is, we shall use σβ+2 (cid:20) (kiki )β+1 N (cid:26) σ2 j + hkβ+1 i(cid:21) + ki + kj − hki(cid:27) . 1 i − kβ+1 hkβ+1i − kβ+1 (C.12) Substituting the adjacency matrix for this expression in the definition of η (Eq. (C.6)), we obtain its expected value as a function of the remaining parameter β : ηβ = hk2i "1 + (σ2 − α2 hkβ+1 i (cid:19)2# , (C.13) hki2 β ρβ ) (cid:18)2 hkβ ihk−1i β ρβ (cid:18) hkβ i hkβ+1i − hk−1i2(cid:19) + α2 where αβ ≡ σ2/σβ+2 and ρβ ≡ hk2(β+1) i − hki2(β+1) . Note that η0 = 1. Fig. C.2 shows the value of ηβ given by Eq. (C.13) against the assor- tativity r for various scale-free networks. Nestedness is seen to grow very fast with increasing disassortativity (decreasing negative r), while in general slightly assortative networks are less nested than neutral ones. However, highly heterogeneous networks (γ → 2) show an increase in ηβ for large positive r . Fig. C.3 shows a plot of nestedness against assortativity for the selection of empirical networks listed in Table C.4. Although these networks are highly disparate as regards size, density, degree distribution, etc., it is apparent from the similarity to Fig. C.2 that the main contribution to ηβ comes indeed from the assortativity. C.5 Bipartite networks Mutualistic networks are usually bipartite: two sets of nodes exist such that all edges are between nodes in one set and those of another. The ones con- sidered in Ref. (Bastolla et al., 2009), for instance, are composed of animals and plants which interact in symbiotic relations of feeding-pollination; these interactions only take place between animals and plants. Let us therefore con- sider a bipartite network and call the sets Γ1 and Γ2 , with n1 and n2 nodes, respectively (n1 + n2 = N ). Using the notation h·ii for averages over set Γi , the total number of edges is hki1n2 = hki2n1 = 1 2 hkiN . Assuming that the network is defined by the configuration ensemble, though with the additional 3Note that C = 1 corresponds to removing the linear term, proportional to ki kj , in Eq. (C.9), and leaving the leading non-linearity, (ki kj )β+1 , as the dominant one. C.5 Bipartite networks 119 γ=2.00 γ=2.00 γ=2.00 γ=2.00 γ=2.00 γ=2.00 γ=2.25 γ=2.25 γ=2.25 γ=2.25 γ=2.25 γ=2.50 γ=2.50 γ=2.50 γ=2.50 γ=2.75 γ=2.75 γ=2.75 γ=3.00 γ=3.00 2.5 2.5 2.5 2.5 2.5 2.5 2 2 2 2 2 2 1.5 1.5 1.5 1.5 1.5 1.5 η η η η η η 1 1 1 1 1 1 0.5 0.5 0.5 0.5 0.5 0.5 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 0 0 0 0 0 0 0.1 0.1 0.1 0.1 0.1 0.1 0.2 0.2 0.2 0.2 0.2 0.2 r r r r r r 0.3 0.3 0.3 0.3 0.3 0.3 0.4 0.4 0.4 0.4 0.4 0.4 0.5 0.5 0.5 0.5 0.5 0.5 0.6 0.6 0.6 0.6 0.6 0.6 Figure C.2: Nestedness against assortativity (as measured by Pearson’s cor- relation coefficient) for scale-free networks as given by Eq. (C.13). hki = 10, N = 1000. 1.75 1.75 1.75 1.5 1.5 1.5 1.25 1.25 1.25 η η η 1 1 1 0.75 0.75 0.75 0.5 0.5 0.5 -0.3 -0.3 -0.3 -0.2 -0.2 -0.2 -0.1 -0.1 -0.1 0 0 0 0.2 0.2 0.2 0.3 0.3 0.3 0.4 0.4 0.4 0.5 0.5 0.5 0.1 0.1 0.1 r r r Figure C.3: Nestedness against assortativity (as measured by Pearson’s cor- relation coefficient) for data on a variety of networks. Blue squares are food webs (Table C.4) and red circles are networks of all other types (Table C.4). constraint of being bipartite, the probability of node l being connected to node i is ǫil = 2 kikl hkiN 120 Chapter C. Nestedness of networks Food web r ν hki N σ/hki Little Rock lake Ythan Estuary (w/p) Stony Stream Canton Creek Skipwith Pond El Verde Caribbean Reef (small) St. Martin Island UK Grassland Chesapeake Bay NE US Shelf Coachella Valley St. Mark’s Estuary -0.343 -0.249 -0.201 -0.196 -0.194 -0.183 -0.172 -0.165 -0.125 -0.123 -0.088 0.043 0.118 1.219 1.323 1.163 1.171 0.891 1.088 1.000 1.071 0.907 0.801 0.971 0.857 0.816 20.4 8.9 14.7 13.5 14.2 18.4 19.7 9.3 2.8 4.1 34.3 14.6 8.5 92 82 109 102 25 155 50 42 61 31 79 29 48 0.73 0.93 0.75 0.69 0.37 0.88 0.49 0.56 0.82 0.60 0.45 0.41 0.55 Table C.1: Food webs appearing in Fig. C.3 (listed from least to most assortative) : r is the assortativity and ν the nestedness. The origins of all data cited in Ref. (Dunne et al., 2004), and kindly provided to us by Jennifer Dunne. ηbip = if they belong to different sets, and zero if they are in the same one. Proceeding as before, we find that the expected value of the nestedness for a bipartite network is N 2 " Xi,j∈Γ1 1 hki2n1 # = kikl 1 klkj kikj Xl∈Γ1 hki1n2 n1hk2 i2 + n2 hk2i1 (C.14) hki1hki2(n1 + n2 )2 . Interestingly, if n1 = n2 , the fact that the network is bipartite has no effect on the nestedness: ηbip = ηconf . 1 kikj Xl∈Γ2 kikl hki1n2 klkj hki2n1 + Xi,j∈Γ2 C.6 Overlapping networks If the adjacency matrix a describes a mutualistic network, the benefit to its being nested resides in a counteraction of the competition matrix c, which C.6 Overlapping networks 121 Network r ν hki N σ/hki Ref. Political blogs Metabolic Political books Adjectives and nouns Dolphins Power grid Neural Jazz musicians Email American football PGP High-energy arXiv Net-science arXiv -0.221 -0.220 -0.138 -0.125 -0.063 0.003 0.005 0.020 0.078 0.133 0.239 0.294 0.462 1.496 1.688 0.996 1.057 0.922 0.834 0.907 0.924 0.923 0.904 0.867 0.533 0.443 22.4 9.0 8.4 7.6 5.1 2.7 5.9 27.6 9.6 10.6 4.6 3.8 3.45 1490 453 104 111 61 4940 306 198 1133 114 10680 8360 1588 1.62 1.87 0.65 0.89 0.58 0.67 0.81 0.63 0.97 0.08 1.77 1.14 1.00 (Adamic and Glance, 2005) (Duch and Arenas, 2005) (Krebs) (Newman, 2006) (Lusseau et al., 2003) (Watts and Strogatz, 1998) (Watts and Strogatz, 1998) (P.Gleiser and Danon, 2003) (Guimer`a et al., 2003) (Girvan and Newman, 2002) (Bogn´a et al., 2004) (Newman, 2001) (Newman, 2006) Table C.2: Empirical networks appearing in Fig. C.3 (listed from least to most assortative) : r is the assortativity and ν the nestedness. All data available on the personal Web pages of ´Alex Arenas, Mark Newman and Duncan Watts. takes into account the extent to which one species is detrimental to another due to predation, sharing of resources, etc. From this point of view, it may be interesting to study to what extent matrices c and a2 overlap (note that both networks have the same nodes, but different edges). Presumably, if ecological networks are assembled in such a way that effective competition is minimised, this overlap should be higher than randomly expected. On the other hand, a certain degree of overlap may also arise from the fact that species interacting symbiotically with the same host are perhaps more than averagely likely to be phylogenetically close and/or phenotypically similar, leading (as Darwin noted) to a higher competition element. In any case, a measure of this overlap is 1 hkicN Xij where h·ic represents an average over the competion network; similarly, h·ia will stand for an average over the mutualistic network. If the two networks cij (a2 )ij , r ≡ (C.15) 122 Chapter C. Nestedness of networks are mutually uncorrelated4 – i.e., if the existence of an edge in one provides no information as to whether there is a corresponding one in the other – we can write cij (C.16) 1 1 hkicN Xij N 2 Xij (a2)ij ≡ runc . r ≃ Using Pij (a2 )ij = hk2 iaN , and assuming that c is normalised so that Pij cij = hkicN , we have5 runc ≃ hk2ia (C.17) , N which only depends on the heterogeneity of the degree distribution of the mutualistic network. Again, it may be useful to consider the overlap normalised to this value, r ≡ = r runc cij (a2)ij . 1 hkic hk2ia Xij This measure will equal unity when there is no statistical relation between the competition matrix and the mutualistic one, but can be expected to be greater if indeed such an overlap were contributing to a reduction in effective competition. (C.18) It has recently been shown that interconnected networks are prone to dan- gerous “cascades of failures” (Buldyrev et al., 2010). It seems that the northen half of Italy was once left temporarily with no electric supply due to failures in the power-grid closing down dependent internet servers, which in turn further disrupted the grid, until many nodes of both networks were rendered dysfunc- tional. If two inter-dependent networks were to coincide perfectly (r = 1), the resilience of the system to node removal would be the same as that of just one network; however, lower overlap leads to increased vulnerability to such cas- cades of failures. Since the extinction of a species can result in its host species also going extinct, such cascades of failures may be a threat to mutualistic systems. In such a case, it would seem that a high overlap r , as defined here, between the competition matrix and the mutualistic one would minimise this possibility. It would be interesting to test this experimentally. 4Note that we are saying nothing of the internal correlations that each network may display. 5The competition matrix will in general be weighted, as could be the mutualistic one; we shall treat both as though they were not, but using weighted networks would only influence results by a normalisation factor. C.7 Discussion C.7 Discussion 123 Whether or not the topological feature here described should be considered a measure of nestedness as it is usually understood in ecology is not clear. What is certain is that interactions between dynamical elements that are mediated by third parties, or common neighbours, can be relevant in a wide variety of settings. We have mentioned the paradigmatical case of ecosystems as well as financial and communications networks. But other examples spring easily to mind. For instance, two excitatory neighbouring neurons might have their mutual effect dampened if they share inhibitory neighbours. Genetic networks are riddled with motifs such that switches activate or inactivate each other indirectly, via common neighbours. As we have shown, there are nontrivial relationships between nestedness, as it is here defined, and other topological features. If it turns out that this network property is indeed relevant for many complex systems, then we hope the null models we have laid out and analysed will prove useful in assessing its functional significance. Appendix D Publications derived from the thesis D.1 Journals and book chapters (the most rel- evant ones marked with an asterisk) 1. * Cluster Reverberation: A mechanism for robust short-term memory without synaptic learning, S. Johnson, J. Marro, and JJ. Torres, submit- ted, arXiv:1007.3122 2. * Enhancing neural-network performance via assortativity, S. de Fran- ciscis, S. Johnson, and J.J. Torres, Physical Review E 83, 036114 (2011) 3. Why are so many networks disassortative? S. Johnson, J.J. Torres, J. Marro, and M.A. Munoz, AIP Conf. Proc. 1332, 249–50 (2011) 4. Shannon entropy and degree-degree correlations in complex networks, S. Johnson, J.J. Torres, J. Marro, and M.A. Munoz, “Nonlinear Systems and Wavelet Analysis”, Ed. R. L´opez-Ruiz, WSEAS Press, pp. 31–35 (2010) 5. * Entropic origin of disassortativity in complex networks, S. Johnson, J.J. Torres, J. Marro, and M.A. Munoz, Physical Review Letters 104, 108702 (2010) 6. * Evolving networks and the development of neural systems, S. Johnson, J. Marro, and J.J. Torres, Journal of Statistical Mechanics (2010) P03003 124 D.2 Abstracts 125 7. Excitable networks: Nonequilibrium criticality and optimum topology, J.J. Torres, S. de Franciscis, S. Johnson, and J. Marro, International Journal of Bifurcation and Chaos 20, 869–875 (2010) 8. Nonequilibrium behavior in neural networks: criticality and optimal per- formance, J.J. Torres, S. Johnson, J.F. Mejias, S. de Franciscis, and J. Marro, “Advances in Cognitive Neurodynamics (II)” Eds. R. Wang and F. Gu, pp 597–603, Springer, 2011, ISBN: 978-90-481-9694-4, Proceed- ings of Second International Conference on Cognitive Neurodynamics (ICCN2009), Hangzhou 15-19 November 2009. 9. Development of neural network structure with biological mechanisms, S. Johnson, J. Marro, J.F. Mejias, and J.J. Torres, Lecture Notes in Com- puter Science 5517, 228–235 (2009) 10. Switching dynamics of neural systems in the presence of multiplicative colored noise, J.F. Mejias, J.J. Torres, S. Johnson, and H.J. Kappen, Lecture Notes in Computer Science 5517, 17–23 (2009) 11. * Nonlinear preferential rewiring in fixed-size networks as a diffusion process, S. Johnson, J.J. Torres, and J. Marro, Physical Review E 79, 050104(R) (2009) 12. * Functional optimization in complex excitable networks, S. Johnson, J.J. Torres, and J. Marro, EPL 83, 46006 (2008) 13. Excitable networks: Non-equilibrium criticality and optimum topology, J.J. Torres, S. de Franciscis, S. Johnson, and J. Marro, “Modelling and Computation on Complex Networks and Related Topics”, Eds. Criado, Gonzalez-Vias, Mancini and Romance. Proceedings of the conference ”Net-Works 2008”, 185–192, ISBN:978-84-691-3819-9. 14. Topology-induced instabilities in neural nets with activity-dependent synapses, S. Johnson, J. Marro, and J. J. Torres, “New Trends and Tools in Com- plex Networks”, Eds. Criado, Pello and Romance. Proceedings of the conference ”Net-Works 2007”, 59–71, ISBN:978-84-690-6890-8. D.2 Abstracts 1. Network topology and dynamical task performance, S. Johnson, J. Marro, and J.J. Torres, AIP Conf. Proc. 1091, 280 (2009) 126 Chapter D. Publications derived from the thesis 2. Constructive chaos in excitable networks with tuneable topologies, S. John- son, J. Marro, and J.J. Torres, XV Congreso de Fsica Estadstica FisEs08, 104 (2008) 3. The effect of topology on neural networks with unstable memories, S. Johnson, J. Marro, and J.J. Torres, AIP Conf. Proc. 887 261 (2006) 4. Relationship between the solar wind and the upper-frequency limit of Saturn Kilometric Radiation, M.Y. Boudjada, P.H.M. Galopeau, H.O. Rucker, A. Lecacheux, W.S. Kurth, D.A. Gurnett, U. Taubencshuss, J.T. Steinberg, S. Johnson, and W. Vollerr, European Geosciences Union (2006) 128 Chapter D. Publications derived from the thesis References H. A. and S. R. Magnetic Domains. Springer, Berlin, 1998. I. Abbott and R. Black. Changes in species composition of floras on islets near perth, western australia. Journal of Biogeography, 7:399–410, 1980. L. Abbott and T. Kepler. From Hodgkin-Huxley to Hopfield, in Statistical mechanics of neural networks. Springer-Verlag, Berlin, 1990. L. Abbott, J. Varela, K. Sen, and S. Nelson. Synaptic depression and cortical gain control. Science, 275:220, 1997. L. Adamic and N. Glance. The political blogosphere and the 2004 US Election. Proceedings of the WWW-2005 Workshop on the Weblogging Ecosystem, 2005. R. Albert and A.-L. Barab´asi. Topology of evolving networks: Local events and universality. Phys. Rev. Lett., 85(24):5234–5237, 2000. R. Albert and A.-L. Barab´asi. Statistical mechanics of complex networks. Rev. Mod. Phys., 74:47–97, 2002. L. Amaral, A. Scala, M. Barth´el`emy, and H. Stanley. Classes of small-world networks. Proc. Natl. Acad. Sci. USA, 97:11149–52, 2000. S. Amari. Characteristics of random nets of analog neuron-like elements. IEEE Trans. Syst. Man. Cybern., 2:643–657, 1972. D. Amit. The Hebbian paradigm reintegraged: Local reverberations as internal representations. Behavioral and Brain Sciences, 18:617–57, 1995. D. J. Amit. Modeling Brain Function: The World of Attractor Neural Net- works. Cambridge University Press, Cambridge, UK, 1989. K. Anand and G. Bianconi. Entropy measures for networks: Toward an infor- mation theory of complex topologies. Phys. Rev. E, 80:045102, 2009. 129 130 REFERENCES L. Antiqueira, F. Rodrigues, and L. Costa. Letter: Modeling connectivity in terms of network activity. Journal of Statistical Mechanics: Theory and Experiment, 9, 2009. A. Arenas, A. D´ıaz-Guilera, and C. P´erez-Vicente. Synchronization reveals topological scales in complex networks. Phys. Rev. Lett., 96:114102, 2006. A. Arenas, A. D´ıaz-Guilera, J. Kurths, Y. Y. Moreno, and C. Zhou. Synchro- nization in complex networks. Phys. Rep., 469:93–153, 2008a. A. Arenas, A. Fern´andez, and S. G. S. A complex network approach to the determination of functional groups in the neural system of c. elegans. Lect. Notes Comp. Sci., 5151:9–18, 2008b. looking back and looking forward. Baddeley and A.D. Working memory: Nature Reviews Neuroscience, 4:829–39, 2003. A. Baddeley. Essentials of Human Memory. Psychology Press, East Sussex, UK, 1999. P. Bak, K. Chen, and C. Tang. A forest-fire model and some thoughts on turbulence. Phys. Lett. A, 147:297, 1990. A. Barab´asi and Z. Oltvai. Network biology: understanding the cell’s func- tional organization. Nature Reviews Genetics, 5:101–3, 2004. A.-L. Barab´asi and R. Albert. Emergence of scaling in random networks. Science, 286:509–512, 1999. M. Barahona and L. Pecora. Synchronization in small-world systems. Phys. Rev. Lett., 89:054101, 2002. O. Barak and M. Tsodyks. Persistent activity in neural networks with dynamic synapses. PLoS Comput. Biol., 3(2):e35, 2007. J. Barrie, J. Freeman, and M. Lenhart. Spatiotemporal analysis of prepyri- form, visual, auditory, and somesthetic surface EEGs in trained rabbits. J. Neurophysiol., 76:520, 1996. U. Bastolla, M. Fortuna, A. Pascual-Garc´ıa, A. Ferrera, B. Luque, and J. Bas- compte. Nature, 458:1018–21, 2009. R. Baxter. Exactly Solved Models in Statistical Mechanics. Academic Press, London, 1982. REFERENCES 131 N. Bertschinger and T. Natschlager. Real-time computation at the edge of chaos in recurrent neural networks. Neural Comp., 16:1413, 2004. G. Bianconi. Mean-field solution of the Ising model on a Barab´asi-Albert network. Phys. Lett. A, 303:166–8, 2002. G. Bianconi. The entropy of randomized network ensembles. EPL, 81:28005, 2008. G. Bianconi. Entropy of network ensembles. Phys. Rev. E, 79(3):036114, 2009. G. Bianconi and A.-L. Barab´asi. Competition and multiscaling in evolving networks. EPL, 54:436–442, 2001. V. Blondel, J.-L. Guillaume, R. Lambiotte, and E. Lefebvre. Fast unfolding of community hierarchies in large networks. J. Stat. Mech., P10008, 2008. S. Boccaletti, V. Latora, Y. Moreno, M. Chavez, and D. Hwang. Complex networks: Structure and dynamics. Physics Reports, 424:175–308, 2006. M. Bogn´a, R. Pastor-Satorras, A. D´ıaz-Guilera, and A. Arenas. Models of social networks based on social distance attachment. Phys. Rev. E, 70: 056122, 2004. M. Bogun´a and R. Pastor-Satorras. Class of correlated random networks with hidden variables. Phys. Rev. E, 68:036112, 2003. M. Bogun´a, F. Papadopoulos, and D. Krioukov. Sustaining the internet with hyperbolic mapping. Nature Communications, 1:62, 2010. B. Bollob´as. Random Graphs. Cambridge University Press, Cambridge, 2001. J. Bonachela, S. de Franciscis, J. Torres, and M. Munoz. Self-organization without conservation: Are neuronal avalanches generically critical? J. Stat. Mech., page P02015, 2010. P. Bonifazi, M. Goldin, M. Picardo, I. Jorquera, A. Cattani, G. Bianconi, A. Represa, Y. Ben-Ari, and R. Cossart. Gabaergic hub neurons orchestrate synchrony in developing hippocampal networks. Science, 326:1419, 2009. M. Brede and S. Sinha. Assortative mixing by degree makes a network more unstable. arXiv:cond-mat/0507710. S. Brush. History of the Lenz-Ising model. Rev. Mod. Phys., 39:883–93, 1967. 132 REFERENCES S. Buldyrev, R. Parshani, G. Paul, H. Stanley, and S. Havlin. Catastrophic cascade of failures in interdependent networks. Nature, 464:1025–8, 2010. E. Bullmore and O. Sporns. Complex brain networks: graph theoretical anal- ysis of structural and functional systems. Nature Reviews Neuroscience, 10: 186–98, 2009. Z. Burda, J. Jurkiewicz, and A. Krzywicki. Statistical mechanics of random graphs. Physica A: Statistical Mechanics and its Applications, 344:56–61, 2004. G. Caldarelli, A. Capocci, P. D. L. Rios, and M. M. oz. Scale-free networks from varying vertex intrinsic fitness. Phys. Rev. Lett., 89:258702, 2002. M. Camperi and X.-J. Wang. A model of visuospatial working memory in prefrontal cortex: recurrent network and cellular bistability. J. Comp. Neu- rosci., 5:383–405, 1998a. M. Camperi and X.-J. Wang. A model of visuospatial working memory in prefrontal cortex: recurrent network and cellular bistability. J. Comp. Neu- rosci., 5:383–405, 1998b. A. Capocci and F. Colaiori. Mixing properties of growing networks and simp- son’s paradox. Phys. Rev. E., 74:026122, 2006. M. Chavez, D.-U. Hwang, A. Amann, H. Hentschel, and S. Boccaletti. Syn- chronization is enhanced in weighted complex networks. Phys. Rev. Lett., 94:218701, 2005. G. Chechik, I. Meilijson, and E. Ruppin. Neuronal regulation: a biologically plausible mechanism for efficient synaptic pruning in development. Neuro- computing, 26:633, 1999. G. Chechik, I. Meilijson, and E. Ruppin. Synaptic pruning in development: A computational account. Neural Computation, in press. D. Chialvo. Critical brain networks. Physica A: Statistical Mechanics and its Applications, 340:756–765, 2004. D. Chialvo. Psychophysics: Are our senses critical? Nature Physics, 2:301–2, 2006. REFERENCES 133 D. Chialvo, P. Balenzuela, and D. Fraiman. The brain: What is critical about it? In . E. P. L. M. Ricciardi, A. Buonocore, editor, Col lective Dynamics: Topics on Competition and Cooperation in the Biosciences, volume 1028 of American Institute of Physics Conference Series, pages 28–45, 2008. A. Compte, C. Constantinidis, J. Tegner, S. Raghavachari, M. Chafee, and et al. Temporally irregular mnemonic persistent activity in prefrontal neurons of monkeys during a delayed response task. J. Neurophysiol., 90:3441–54, 2003. R. Conrad. Acoustic confusion in immediate memory. B. J. Psychol., 55: 75–84, 1964a. R. Conrad. Information and acoustic confusion and memory span. B. J. Psychol., 55:429–32, 1964b. J. Cortes, P. Garrido, H. Kappen, J. Marro, C. Morilla, D. Navidad, and J. Torres. Algorithms for identification and categorization. AIP Conf. Proc., 779:178, 2005. J. Cortes, J. Torres, J. Marro, P. Garrido, and H. Kappen. Effects of fast presynaptic noise in attractor neural networks. Neural Comp., 18:614, 2006. N. Cowan. On short and long auditory stores. Psychological Bul letin, 96: 341–70, 1984. V. de Assis and M. Copelli. Dynamic range of hypercubic stochastic excitable media. Phys. Rev. E, 77:011923, 2008. S. de Franciscis, S. Johnson, and J. Torres. Enhancing neural-network perfor- mance via assortativity. Phys. Rev. E, in press, 2011. D. J. de S. Price. Networks of scientic papers. Science, 149:510–5, 1965. I. Der´enyi, I. Farkas, G. Palla, and T. Vicsek. Topological phase transitions of random networks. Physica A: Statistical Mechanics and its Applications, 334:583–590, 2004. D. Dominguez, M. Gonz´alez, E. Serrano, and F. Rodr´ıguez. Structured infor- mation in small-world neural networks. Phys. Rev. E, 79:021909, 2009. V. Dom´ınguez-Chibet´ın, S. Johnson, and M. Munoz. Hidden interactions in food-webs: Unearthing the missing links. in preparation, 2011. 134 REFERENCES L. Donetti and M. A. Munoz. Detecting network communities: a new sys- tematic and powerful algorithm. J. Stat. Mech.: Theor. Exp., page P10012, 2004. S. Dorogovtsev and J. Mendes. Evolution of networks: From biological nets to the Internet and WWW. Oxford Univ. Press, Oxford, 2003. S. Dorogovtsev, A. Ferreira, A. Goltsev, and J. Mendes. Zero pearson coeffi- cient for strongly correlated growing trees. Phys. Rev. E, 81:031135, 2005. J. Duch and A. Arenas. Community identification using Extremal Optimiza- tion. Phys. Rev. E, 72:027104, 2005. J. Dunne, R. Williams, and N. Martinez. Network structure and robustness of marine food webs. Mar. Ecol. Prog. Ser., 273:291–302, 2004. D. Durstewitz, J. Seamans, and T. S. and. Neurocomputational models of working memory. Nature Neuroscience, 3:1184–91, 2000. V. Egu´ıluz, D. Chialvo, G. Cecchi, M. Baliki, and A. Apkarian. Scale-free brain functional networks. Phys. Rev. Lett., 94:018102, 2005. P. Erdos and A. R´enyi. On random graphs I. Publ. Math. Debrecen, 6:290–7, 1959. L. Euler. Solutio problematis ad geometriam situs pertinentis. Acad. Sci. U. Petrop., 8:128–40, 1736. I. Farkas, I. Der´enyi, G. Palla, and T. Vicsek. Lect. Notes in Phys., 650:163, 2004. E. Frank. Synapse elimination: For nerves it’s all or nothing. Science, 275: 324–325, 1997. A. Fronczak and P. Fronczak. Networks with given two-point correlations: hidden correlations from degree correlations. Phys. Rev. E, 74:026121, 2006. G. G. Bianconi and A.-L. Barab´asi. Bose-Einstein condensation in complex networks. Phys. Rev. Lett., 86:5632–5635, 2001. M. Garey and D. Johnson. Computers and Intractability: A Guide to the Theory of NP-Completeness. W.H. Freeman and Company, USA, 1979. REFERENCES 135 E. Gilbert. Random graphs. Annals of Mathematical Statistics, 30:1141–4, 1959. M. Girvan and M. Newman. Community structure in social and biological networks. Proc. Natl. Acad. Sci. USA, 99:7821–6, 2002. J. G´omez-Gardenes, Y. Moreno, and A. Arenas. Paths to synchronization on complex networks. Phys. Rev. Lett., 98:034101, 2007. A. Gruart, M. Munoz, and J. Delgado-Garc´ıa. Involvement of the ca3-ca1 synapse in the acquisition of associative learning in behaving mice. J. Neu- rosci., 26:1077–87, 2006. R. Guimer`a, L. Danon, A. D´ıaz-Guilera, F. Giralt, and A. Arenas. Self-similar community structure in a network of human interactions. Phys. Rev. E, 68: 065103(R), 2003. D. Hebb. The Organization of Behavior. Wiley, New York, 1949. S. Hilfiker, V. Pieribone, H.-T. Kao, A. Czernik, G. Augustine, and P. Green- gard. Synapsins as regulators of neurotransmitter release. Phil. Trans. R. Soc. London B, 354:269–79, 1999. A. Hodgkin and A. Huxley. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol., 117: 500–44, 1952. H. Holcman and M. Tsodyks. The emergence of Up and Down states in cortical networks. PLoS Comput. Biol., 2:e23, 2006. P. Holme and J. Zhao. Exploring the assortativity-clustering space of a net- work’s degree sequence. Phys. Rev. E, 75:046111, 2007. J. Hopfield. Neural networks and physical systems with emergent collective computational abilities. Proc. Natl. Acad. Sci. USA, 79:2554–8, 1982. P. Hurtado, J. Marro, and P. Garrido. Demagnetization via nucleation of the nonequilibrium metastable phase in a model of disorder. J. Stat. Phys., 133: 29–58, 2008. P. R. Huttenlocher and A. S. Dabholkar. Regional differences in the synap- togensis in human cerebral cortex. The Journal of Comparative Neurology, 387:167–178, 1997. 136 REFERENCES E. M. Izhikevich. Dynamical Systems in Neuroscience: The Geometry of Ex- citability and Bursting. MIT Press, Cambridge MA, 2007. E. Jaynes. Information theory and statistical mechanics. Phys. Rev., 106(4): 620–630, 1957. S. Johnson and M. Munoz. Nestedness of networks: Null models and topolog- ical effects. in preparation. S. Johnson, J. Torres, and J. Marro. Functional optimization in complex excitable networks. EPL, 83:46006, 2008. S. Johnson, J. Marro, and J. Torres. Development of neural network structure with biological mechanisms. Lecture Notes in Computer Science, 5517:228– 35, 2009a. S. Johnson, J. Torres, and J. Marro. Nonlinear preferential rewiring in fixed- size networks as a diffusion process. Phys. Rev. E, 79:050104, 2009b. S. Johnson, J. Marro, and J. Torres. Evolving networks and the development of neural systems. J. Stat. Mech., P03003, 2010a. S. Johnson, J. Torres, J. Marro, and M. Munoz. Entropic origin of disassorta- tivity in complex networks. Physical Review Letters, 104:108702, 2010b. S. Johnson, J. Marro, and J. Torres. Cluster Reverberation: A mecha- nism for robust short-term memory without synaptic learning. submitted. arXiv:1007.3122, 2011. M. Kaiser and C. Hilgetag. Spatial growth of real-world networks. Phys. Rev. E, 69:036103, 2004. M. Kaiser, R. Martin, P. Andras, and M. Young. Simulation of robustness against lesions of cortical networks. Eur. J. Neurosci., 25:3185–92, 2007. O. Kinouchi and M. Copelli. Optimal dynamical range of excitable networks at criticality. Nature Physics, 2(5):348–351, 2006. A. Klintsova and W. Greenough. Synaptic plasticity in cortical systems. Cur- rent Opinion in Neurobiology, 9:203, 1999. H. Korn and P. Faure. Is there chaos in the brain? II. Experimental evidence and related models. C. R. Biol., 326:787, 2003. REFERENCES 137 G. Kossinets and D. Watts. Science, 311:88–90, 2006. P. L. Krapivsky, S. Redner, and F. Leyvraz. Connectivity of growing random networks. Phys. Rev. Lett., 85:4629–4632, 2000. V. Krebs. unpublished, http://www.orgnet.com/. A. Kwag and O. Paulsen. The timing of external input controls the sign of plasticity at local synapses. Nat Neurosci., 12:1219–21, 2009. J. D. L. Danon, A. D´ıaz-Guilera, and A. Arenas. Comparing community structure identification. J. Stat. Mech., P09008, 2005. K. Lee, F. Schottler, M. Oliver, and G. Lynch. Brief bursts of high-frequency stimulation produce two types of structural change in rat hippocampus. J. Neurophysiol., 44:247, 1980. Levine and R.D. Molecular Reaction Dynamics. Cambridge University Press, Cambridge, 2005. B. Lindner, J. Garc´ıa-Ojalvo, A. Neiman, and L. Schimansky-Geier. Effects of noise in excitable systems. Phys. Rep., 392:321, 2004. D. Lusseau, K. Schneider, O. J. Boisseau, P. Haase, E. Slooten, and S. Daw- son. The bottlenose dolphin community of doubtful sound features a large proportion of long-lasting associations. Behavioral Ecology and Sociobiology, 54:396–405, 2003. P. J. Magistretti. Neuroscience. low-cost travel in neurons. Science, 325:1349– 51, 2009. R. Malenka and R. Nicoll. Long-term potentiation – a decade of progress? Science, 285:1870–4, 1999. Marcus and G.F. The Algebraic Mind: Integrating Connectionism and Cogni- tive Science. MIT Press, Cambrigde and MA, 2001. J. Marro and R. Dickman. Nonequilibrium Phase Transitions in Lattice Models. Cambridge. J. Marro, J. Torres, and J. Cortes. Complex behavior in a network with time- dependent connections and silent nodes. J. Stat. Mech., page P02017, 2008. 138 REFERENCES S. Maslov, K. Sneppen, and A. Zaliznyak. Pattern detection in complex net- works: correlation profile of the internet. Physica A, 333:529–40, 2004. R. May. Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton, 1973. W. McCulloch and W. Pitts. A logical calculus of the ideas immanent in nervous activity. Bul letin of Mathematical Biophysics, 7:115–33, 1943. J. Mejias and J. Torres. Maximum memory capacity on neural networks with short-term synaptic depression and facilitation. Neural Comput., 21:851–71, 2009. J. Mejias, H. Kappen, and J. Torres. Irregular dynamics in up and down cortical states. PLoS ONE, 5 (11):e13651, 2010. E. Meron. Pattern formation in excitable media. Phys. Rep., 218:1, 1992. S. Milgram. The Small World problem. Psychology Today, 1:61–67, 1967. Y. Miyashita. Neuronal correlate of visual associative long-term memory in the primate temporal cortex. Nature, 335:817–20, 1988. G. Mongillo, O. Barak, and M. Tsodyks. Synaptic theory of working memory. Science, 319:1543–6, 2008. R. Morgan and I. Soltesz. Nonrandom connectivity of the epileptic dentate gyrus predicts a ma jor role for neuronal hubs in seizures. Proc. Nat. Acad. Sci. USA, 105:6179, 2008. A. Motter, C. Zhou, and J. Kurths. Network synchronization, diffusion, and the paradox of heterogeneity. Phys. Rev. E, 71:016116, 2005. M. Munoz, R. Juhasz, C. Castellano, and G. Odor. Griffiths phases on complex networks. Phys. Rev. Lett., 105:128701, 2010. M. Newman. The structure of scientific collaboration networks. Proc. Natl. Acad. Sci. USA, 98:404–9, 2001. M. Newman. Assortative mixing in networks. Phys. Rev. Lett., 89:208701, 2002. M. Newman. Mixing patterns in networks. Phys. Rev. E, 67:026126, 2003a. REFERENCES 139 M. Newman. Random graphs as models of networks, in Handbook of Graphs and Networks, S. Bornholdt and H.G. Schuster (eds.). Wiley-VCH, Berlin, 2003b. M. Newman. The structure and function of complex networks. SIAM Review, 45:167–256, 2003c. M. Newman. Power laws, Pareto distributions and Zipf ’s law. Contemporary Physics, 46:323, 2005. M. Newman and J. Park. Why social networks are different from other types of networks. Phys. Rev. E, 68:036122, 2003. M. E. J. Newman. Finding community structure in networks using the eigen- vectors of matrices. Preprint physics/0605087, 2006. T. Nishikawa, A. Motter, Y.-C. Lai, and F. Hoppensteadt. Heterogeneity in oscillator networks: Are smaller worlds easier to synchronize? Phys. Rev. Lett., 91:014101, 2003. L. Onsager. Crystal statistics. I. A two-dimensional model with an order- disorder transition. Phys. Rev., (2) 65:117–49, 1944. L. Pantic, J. Torres, H. Kappen, and S. Gielen. Associative memory with dynamic synapses. Neural Comp., 14:2903, 2002. J. Park and M. Newman. Origin of degree correlations in the internet and other networks. Phys. Rev. E, 68:026112, 2003. J. Park and M. E. J. Newman. Statistical mechanics of networks. Phys. Rev. E, 70:066117, 2004. K. Park, Y.-C. Lai, and N. Ye. Self-organized scale-free networks. Phys. Rev. E, 72:026131, 2005. R. Pastor-Satorras and A. Vespignani. Evolution and structure of the Internet: A statistical physics approach. Cambridge Univ. Press, Cambridge, 2004. R. Pastor-Satorras, A. V´azquez, and A. Vespignani. Dynamical and correlation properties of the internet. Phys. Rev. Lett., 87:258701, 2001. T. Peixoto. Redundancy and error resilience in boolean networks. Phys. Rev. Lett., 104:048701, 2010. 140 REFERENCES T. Pereira. Hub synchronization in scale-free networks. Phys. Rev. E, 82: 036201, 2010. P. Peretto. An Introduction to the Modeling of Neural Networks. Cambridge Univ. Press, Cambridge, 1992. P.Gleiser and L. Danon. Community structure in Jazz. Adv. Complex Syst., 6:565, 2003. J. Poncela, J. Gmez-Gardees, L. Flor´ıa, A. S´anchez, and Y. Moreno. Complex cooperative networks from evolutionary preferential attachment. PLoS ONE, 3(6):e2449, 2008. S. Ree. Generation of scale-free networks using a simple preferential-rewiring dynamics. Physica A: Statistical Mechanics and its Applications, 376:692– 698, 2007. J. Reichardt and S. Bornholdt. Statistical mechanics of community detection. Phys. Rev. E, 74:016110, 2006. A. Rodr´ıguez-Moreno and O. Paulsen. Spike timing-dependent long-term de- pression requires presynaptic nmda receptors. Nat Neurosci., 11:744–5, 2008. M. D. Roo, P. Klauser, P. Mendez, L. Poglia, and D. Muller. Activity- dependent PSD formation and stabilization of newly formed spines in hip- pocampal slice cultures. Cerebral Cortex, 18:151–161, 2008. Y. Roudi and P. Latham. A balanced memory network. PLoS Comput. Biol., 3. A. Rozenfeld, R. Cohen, D. ben Avraham, and S. Havlin. Scale-free networks on lattices. Phys. Rev. Lett., 89:218701, 2002. K. Sakai and Y. Miyashita. Neuronal organization for the long-term memory of paired associates. Nature, 354:152–5, 1991. I. Sendina-Nadal, J. Buld´u, I. Leyva, and S. Boccaletti. Phase locking induces scale-free topologies in networks of coupled oscillators. PLoS ONE, 3:e2644, 2008. D. Sherrington and S. Kirkpatrick. Solvable model of a spin-glass. Phys. Rev. Lett., 35:1792–6, 1975. REFERENCES 141 S. Sikstrom. Forgetting curves: implications for connectionist models. Cogni- tive Psychology, 45:95–152, 2002. H. A. Simon. On a class of skew distribution functions. Biometrika, 42:425–40, 1955. B. Soderberg. General formalism for inhomogeneous random graphs. Phys. Rev. E, 66:066121, 2002. G. Sperling. The information available in brief visual persentation. Psycho- logical Monographs: General and Applied, 74:1–30, 1960. O. Sporns, D. R. Chialvo, M. Kaiser, and C. C. Hilgetag. Organization, de- velopment and function of complex brain networks. Trends in Cognitive Sciences, 8:418–425, 2004. K. Suchecki, V. Egu´ıluz, and M. S. Miguel. Conservation laws for the voter model in complex networks. EPL, 69:228–34, 2005. G. Suel, J. Garcia-Ojalvo, L. Liberman, and M. Elowitz. An excitable gene regulatory circuit induces transient cellular differentiation. Nature, 440:545– 50, 2006. G. Sugihara and H. Ye. Complex systems: Cooperative network dynamics. Nature, 458:979–80, 2009. E. Tarnow. Short term memory may be the depletion of the readily releasable pool of presynaptic neurotransmitter vesicles. Cognitive Neurodynamics, 3: 263–9, 2008. J.-N. Teramae and T. Fukai. A cellular mechanism for graded persistent ac- tivity in a model neuron and its implications for working memory. J. Comp. Neurosci., 18:105–21, 2005. S. Thurner, F. Kyriakopoulos, and C. Tsallis. Phys. Rev. E., 76:036111, 2007. J. Torres and P. Varona. Modeling Biological Neural Networks, in Handbook of Natural Computing, ed. G. Rozenberg et al. Springer Verlag, Berlin, Ger- many, 2010. J. Torres, M. Munoz, J. Marro, and P. L. Garrido. Influence of topology on the performance of a neural network. Neurocomputing, 58-60:229–234, 2004. 142 REFERENCES J. Torres, J. Cortes, J. Marro, and H. Kappen. Competition between synaptic depression and facilitation in attractor neural networks. Neural Computa- tion, 19:2739–55, 2007. J. Torres, J. Marro, J. M. Cortes, and B. Wemmenhove. Instabilities in at- tractor networks with fast synaptic fluctuations and partial updating of the neurons activity. Neural Nets., 21:1272–77, 2008. J. Torres, J. Marro, and S. de Franciscis. Chaos in heterogeneous networks with temporally inert nodes. Int. J. Bifur. Chaos, 19:677–86, 2009. M. Tsodyks, K. Pawelzik, and H. Markram. Neural networks with dynamic synapses. Neural Comp., 10:821–35, 1998. N. van Kampen. Stochastic Processes in Physics and Chemistry. Elsevier Science, Amsterdam, 1992. F. Vazquez, V. Egu´ıluz, and M. S. Miguel. Generic absorbing transition in coevolution dynamics. Phys. Rev. Lett., 100:108702, 2008. T. Vogels, K. Ra jan, and L. Abbott. Neural network dynamics. Annu. Rev. Neurosci., 28:357–76, 2005. X.-J. Wang. Synaptic reverbaration underlying mnemonic presistent activity. TRENDS in Neurosci., 24:455–63, 2001. D. J. Watts and S. H. Strogatz. Collective dynamics of ’small-world’ networks. Nature, 393:440–442, 1998. J. White, E. Southgate, J. N. Thomson, and S. Brenner. The structure of the nervous system of the nematode c. elegans. Philosophical transactions Royal Society London, 314:1–340, 1986. J. Wixted and E. Ebbesen. On the form of forgetting. Psychological Science, 2:409–15, 1991. J. Wixted and E. Ebbesen. Genuine power curves in forgetting: A quantitative analysis of individual sub ject forgetting functions. Memory & Cognition, 25: 731–9, 1997. S. R. y Ca jal. Histology of the Nervous System of Man and Vertebrates. Oxford University Press, Oxford, 1995. REFERENCES 143 C. Zhou, A. Motter, and J. Kurths. Universality in the synchronization of weighted random networks. Phys. Rev. Lett., 96:034101, 2006a. C. Zhou, L. Zemanov´a, G. Zamora, C. Hilgetag, and J. Kurths. Hierarchical organization unveiled by functional connectivity in complex brain networks. Phys. Rev. Lett., 97:238103, 2006b.
1806.06809
2
1806
2018-07-05T16:48:00
Drifting perceptual patterns suggest prediction errors fusion rather than hypothesis selection: replicating the rubber-hand illusion on a robot
[ "q-bio.NC" ]
Humans can experience fake body parts as theirs just by simple visuo-tactile synchronous stimulation. This body-illusion is accompanied by a drift in the perception of the real limb towards the fake limb, suggesting an update of body estimation resulting from stimulation. This work compares body limb drifting patterns of human participants, in a rubber hand illusion experiment, with the end-effector estimation displacement of a multisensory robotic arm enabled with predictive processing perception. Results show similar drifting patterns in both human and robot experiments, and they also suggest that the perceptual drift is due to prediction error fusion, rather than hypothesis selection. We present body inference through prediction error minimization as one single process that unites predictive coding and causal inference and that it is responsible for the effects in perception when we are subjected to intermodal sensory perturbations.
q-bio.NC
q-bio
Drifting perceptual patterns suggest prediction errors fusion rather than hypothesis selection: replicating the rubber-hand illusion on a robot Nina-Alisa Hinz†, Pablo Lanillos‡∗, Hermann Mueller†, Gordon Cheng‡ † Experimental Neuro-Cognitive Psychology, Ludwig-Maximilians-Universitat, Munchen, Germany ‡ Institute for Cognitive Systems, Technische Universitat Munchen, Munchen, Germany ∗Email: [email protected] 8 1 0 2 l u J 5 ] . C N o i b - q [ 2 v 9 0 8 6 0 . 6 0 8 1 : v i X r a Abstract-Humans can experience fake body parts as theirs just by simple visuo-tactile synchronous stimulation. This body- illusion is accompanied by a drift in the perception of the real limb towards the fake limb, suggesting an update of body estimation resulting from stimulation. This work compares body limb drifting patterns of human participants, in a rubber hand illusion experiment, with the end-effector estimation displace- ment of a multisensory robotic arm enabled with predictive processing perception. Results show similar drifting patterns in both human and robot experiments, and they also suggest that the perceptual drift is due to prediction error fusion, rather than hypothesis selection. We present body inference through prediction error minimization as one single process that unites predictive coding and causal inference and that it is responsible for the effects in perception when we are subjected to intermodal sensory perturbations. Index Terms-Sensorimotor self, rubber-hand illusion, predic- tive coding, robotics I. INTRODUCTION Distinguishing between our own body and that of others is fundamental for our understanding of the self. By learning the relationship between sensory and motor information and integrating them into a common percept, we gradually develop predictors about our body and its interaction with the world [1]. This body learning is assumed to be one of the major processes underlying embodiment. Body-ownership illusions, like the rubber hand illusion [2], are the most widely used methodology to reveal information about the underlying mech- anisms, helping us to understand how the sensorimotor self is computed. Empirical evidence has shown that embodiment is flexible, adaptable, driven by bottom-up and top-down modulations and sensitive to illusions. We replicated the passive rubber hand illusion on a multi- sensory robot and compared it with human participants, there- fore gaining insight into the perceptive contribution to self- computation. Enabling a robot with human-like self-perception [3] is important for: i) improving the machine adaptability has by been supported This work SELFCEPTION project (www.selfception.eu) European Union Horizon 2020 Programme (MSCA- IF-2016) under grant agreement n. 741941 and the ENB Master Program in Neuro-Cognitive Psychology at Ludwig-Maximilians Universitat. Video to this paper: http://web.ics.ei.tum.de/~pablo/rubberICDL2018PL.mp4 Accepted for publication at 2018 IEEE International Conference on Devel- opment and Learning and Epigenetic Robotics and providing safe human-robot interaction, and ii) testing computational models for embodiment and obtaining some clues about the real mechanisms. Although some computa- tional models have already been proposed for body-ownership, agency and body-illusions, the majority of them are restricted to the conceptual level or simplistic simulation [4]. Examining real robot data and using body illusions as a benchmark for testing the underlying mechanism enriches the evaluation considerably. To the best of our knowledge, this is the first study replicating the rubber hand illusion on an artificial agent and comparing it to human data. the robot We showed that when inferring the robot body location through prediction error minimization [5], limb drifting patterns are similar to those observed in human par- ticipants. Human and robot data suggest that the perception of the real hand and the rubber hand location drifts to a common location between both hands. This supports the idea that, instead of selecting one of two hypotheses (common cause for stimulation vs. different causes) [6], visual and proprioceptive information sources are merged generating an effect similar to averaging both hypotheses [7]. The remainder of this paper is as follows: Sec. II describes current rubber-hand illusion findings and its neural basis; Sec. III defines the computational model designed for the robot; Sec. IV describes the experimental setup for both human participants and the robot; Sec. V presents the comparative analysis of the drifting patterns; Finally, Sec. VI discusses body estimation within the prediction error paradigm as the potential cause of the perceptual displacement. II. BACKGROUND A. Rubber-hand illusion Botvinick and Cohen [10] demonstrated that humans can embody a rubber hand only by means of synchronous visuo- tactile stimulation of the rubber hand and their hidden real hand. This was measured using a questionnaire about the illusion, but also by proprioceptive localization of the partici- pant's real hand. After experiencing the illusion, the perception of their own hand's position had drifted towards that of the rubber hand. Since then, multiple studies have replicated the illusion under different conditions (for a review, see [11]). Collectively, these studies show that top-down expectations found an increased proprioceptive drift for a larger distance, the relative amount of drift (i.e. corrected for distance) did not differ between the small and large distances. In [17], they replicated this result, as long as the fake hand was near the body. If the real hand was closer to body midline than the fake hand, increasing distance between both hands decreased the proprioceptive drift. In the present study, we systematically varied the distance between both hands. The real hand, however stayed in the same position for all conditions and only the fake hand had a varying distance from the real hand in anatomically plausible positions. In [8], where the distance between both hands was varied by displacing the fake hand in relation to the real hand, the fake hand was also increasingly rotated with increasing distance. Rotational differences, nevertheless, may influence the illusion [20], probably confounding the results of [8]. In the current study, we systematically examined the influence of the distance between the rubber and the real hand on proprioceptive and visual drift. This provided the basis for validating the computational model proposed in Sec. III and comparing the drift of the body estimation in different distances between the robot and humans. B. Body illusions in the brain A seminal contribution to possible neuronal mechanisms underlying the rubber hand illusion came from [21], who discovered parietal neurons in the primate brain coding for the position of the real arm and a plosturally plausible fake monkey arm. Several fMRI studies looked into the neural correlates for body-ownership illusions in humans (see [20] for a review). Three areas were consistently found activated during the rubber hand illusion: posterior parietal cortex (including intra-parietal cortex and temporo-parietal junction), premotor cortex and lateral cerebellum. The cerebellum is assumed to compute the temporal relationship between visual and tactile signals, thus playing a role in the integration of visual, tactile and proprioceptive body-related signals [22], [23]. The premotor and intra-parietal cortex are multisensory areas, also integrating visual, tactile and proprioceptive signals present during the rubber hand illusion [24]. In [20], they differentiated the role of posterior parietal cortex, being responsible for the recalibration of visual and tactile coordinate systems into a common reference frame, and the role of premotor cortex, being responsible for the integration of signals in this common, hand-centered reference frame. Although it is known that these areas participate in evoking the rubber hand illusion, little is known about the underlying computations [25]. In [26], they used dynamic causal modeling during the rubber hand illusion to confirm to some extent that, during the illusion, visual in- formation is weighted more than proprioceptive information - which would be congruent with predictive coding models. Dur- ing the illusion the intrinsic connectivity of the somatosensory cortex was reduced, indicative of less somatosensory precision, while the extrinsic connectivity between the lateral occipital complex and premotor cortex was increased, indicative of (a) Illusion depending on distance (b) Proprio. drift depending on time Fig. 1. (a) Quality of the illusion as a function of the distance between hands: the higher the better. (b) Intensity of the proprioceptive drift depending on the visuo-tactile stimulation time. Adapted from [8] and [9] respectively. about the physical appearance of a human hand, resulting from abstract internal body models, and bottom-up sensory information, especially spatiotemporal congruence of the stim- ulation and distance between the hands, influence embodiment of the fake hand [9]. In [12], they were even able to induce body-ownership on a robotic arm. The common assumption is that the spatial representations of both hands are merged, as a result of minimizing the error between predicted sensory outcomes from seeing the stimulation of the rubber hand and the actual sensory outcome of feeling the stimulation of one's own hand [13]. Recently, [7] undermined this by showing that not only the perception of the real hand's position is drifted towards the rubber hand (proprioceptive drift), but the one of the rubber hand is drifted towards the real hand as well (visual drift), i.e. to a common location between both hands. In [6], they proposed a Bayesian Causal Inference Model for this prediction error minimization, considering visual, tactile and proprioceptive information, weighted according to their precision. In combination with the prior probability of assum- ing a common cause or different causes for the conflicting multi-sensory information, the posterior probability of each hypothesis is computed. A common cause, i.e. ownership of the rubber hand, is assumed if the posterior probability exceeds a certain threshold. This binarity of the illusion, however, is at variance with findings of [9], demonstrating a continuous proprioceptive drift of the stimulated hand. The proprioceptive drift was shown to increase exponentially during the first minute of stimulation and increasing further over the following four minutes (Fig. 1(b)). Although the reported onset of the illusion ranged from 11 seconds [14] to 23 seconds, with 90 percent of subjects experiencing it within the first minute of stimulation [15], the ongoing drifting suggests a continuous, rather than a discrete mechanism, being involved in embodiment. The proposed computational models in the literature of body-ownership illusions need further verification from exper- imental data. Several studies showed reduced illusion scores for larger, as compared to smaller, distances between the real and the rubber hand [8], [16]–[19] (Fig. 1(a)), though only few studies measured the proprioceptive drift in dependence on the distance between the two hands [16], [17]. While [16] more attention to visual input. Further functional evidence for the proposed computations is needed. III. COMPUTATIONAL MODEL (cid:88) i x = (si − gi(x)) (cid:124) (cid:123)(cid:122) σsi error expected sensation (cid:125) i(x) − sx − µx (cid:124) (cid:123)(cid:122) (cid:125) g(cid:48) σx error prior (3) We assume that all sensations / features follow a Gaussian dis- tribution with (linear or non-linear) mean gi(x) and variance σsi. The forward models learned should be differentiable with respect to the body configuration (g(cid:48) By rewriting the prediction error as e = s − g(x) and defining µx as the prior belief about the body configuration, the dynamics of the body perception model are described by (see Appendix for derivation and [27] for the detailed algorithm): i(x) = ∂gi(x)/∂x). v(x) + evtg(cid:48) vt(x) x = −ex + ep + evg(cid:48) ex = sx − µx − σxex ep = sp − x − σpep ev = sv − gv(x) − σvev evt = svt − gvt(x) − σvtevt µx = µx + λex (4) where λ is the learning ratio parameter that specifies how fast the prior of the body configuration µx is adjusted to the prediction error. The visual forward function gv and its derivative are cal- culated using Gaussian process estimation (see Sec. IV). The visuo-tactile generative function is computed by means of a hand-crafted likelihood, which uses the visual ov and tactile st stimulation information (temporal hs and spatial ht), and the expected position of the hand gv(x): gt(x) = st · hs · ht = sta1e−b1 (cid:80)(gv(x)−ov)2 · a2e−b2δ2 (5) where a1, b1, a2, b2 are parameters that shape the likelihood of the spatial plausibility and have been tuned in accordance with the data acquired from human participants; δ is the level of synchrony of the events (e.g. timing difference between the visual and the tactile event); and ov is the other agent end- effector location in the visual field. IV. EXPERIMENTAL SET-UP Fig. 2. Rubber hand illusion modelled as a body estimation problem solved using prediction error minimization. Visual features of the rubber hand are incorporated when there is synchronous visuo-tactile stimulation, though it is constrained by the prior belief and the expected location of the hand according to the generative visual forward model and the estimated body configuration. We formalized the rubber hand illusion as a body estimation problem under the predictive processing framework. The core idea behind this is that all features and sensory modalities are contributing to refine body estimation through the min- imization of the errors between sensations and predictions [27]. During synchronous visuo-tactile stimulation, the most plausible body configuration is perturbed due to the merging of visual and proprioceptive information. This is coherent with the drift of both the real hand and the rubber hand as the participants are just pointing to the prediction of their hand according to the current body configuration. Figure 2 shows how sensory modalities or features are contributing to the estimation of the participants limb. We define x as the latent space variable that expresses the body configuration. We model the problem as inferring the most plausible body configuration x given the sensation likelihood and the prior: P (xs) = p(sx)p(x). We further define sp, sv, svt as the proprioceptive, visual and visuo-tactile sensation respectively. Assuming independence of the different sources of information we get: P (xs) = p(spx)p(svx)p(svtx)p(x) (1) A. Participants selection The perception or estimation of the body is then solved by learning an approximation of the forward model for each feature or modality s = g(x) and minimizing a lower bound on the KL-divergence known as negative free energy F [5], [28]. log P (xs) = −F = log p(six) + log p(x) (2) (cid:88) i We obtain the estimated value of the latent variables through gradient descent minimization x = ∂F x : 20 volunteers (mean age: 25, 75 % female) took part in the experiment. They received 8 euros per hour in compensation. All participants were right-handed, had no disability of per- ceiving touch on their right hand, did not wear nail polish and did not have any special visual features (e.g. scars / tattoos) on their right hand. They had no neurological or psychological disorders, as indicated by self-report, and normal or corrected- to-normal vision. None of them had experienced the rubber hand illusion before. All participants gave informed consent prior to the experiment. (a) Human (b) Robot Fig. 3. Experimental setups used in the current study. (a) Rubber-hand illusion in different distance conditions and (b) artificial version. B. Humans experiment details We performed the rubber-hand illusion experiment, focusing on the proprioceptive drift of the real and the visual drift of the rubber hand, as a function of the distance between both hands. The experiment, depicted in Fig. 3, comprised six conditions, each with a different distance between the real hand and the rubber hand. The participant's real right hand was placed in a box, with the index finger 20cm away from the participants body midline. The rubber hand was placed with its index finger 15cm, 20cm, 25cm, 30cm, 35cm or 40cm away from the participants real right hand (5cm, 0cm, -5cm, -10cm, -15cm or -20cm away from the participants body midline respectively). Participants sat in front of a wooden box, placing their hands near the outer sides of the box. They wore a rain cape covering their body and arms. In one of the arms, a rubber hand was placed such that it seemed coherent with the body. With their left hand they held a computer mouse. Each trial consisted of four phases: localization of the real hand, localization of the rubber hand, the stimulation phase and post-stimulation localization of both hands. 1) First, we covered the box with a wooden top and a blanket above it, so that no visual cues could be used. Participants had to indicate where they currently perceived the location of the index finger of their right hand, pointing with the mouse on a horizontal line presented on the screen. The line did cover the whole length of the box. 2) After fixating the rubber hand for one minute, we again covered the box and the same task was repeated for the rubber hand. 3) The box was remodeled, removing the cover and intro- ducing a vertical board next to the participant's right hand so that it was not visible to the participant (Fig. 3(a)). Then the experimenter began stimulating the rub- ber and the real hand synchronously with two similar paintbrushes, starting at the index finger, continuing to the little finger and then starting at the index finger again, with one brush stroke each about two seconds. 4) Subsequently, participants were again asked to indicate where they perceived the index finger of the real or the rubber hand, starting with the real or the rubber hand in randomized fashion. The box was covered during the localization. 5) At the end of each trial, participants were asked to answer ten questions related to the illusion adapted from [29], presented randomized on the screen, using a con- tinuous scale from -100 (indicating strong disagreement) to 100 (indicating strong agreement). For the localization trials, a horizontal line was presented on the screen opposite to the box, with the screen having the same size as the box. The localization trials were repeated ten times to account for high variance. The proprioceptive drift and visual drift were calculated by subtracting the average of the first localization phase from the second localization phase for the real hand and the rubber hand separately. The illusion index was calculated by subtracting the average response to control statements S4-S10 from the average response to illusion statements S1-S3 [30]. Between all phases participants were blindfolded, so they did not observe the remodeling, which might potentially have impeded the illusion. C. Robot experiment details We tested the model on the multisensory UR-5 arm of robot TOMM [31], as depicted in Fig. 3(b). The proprioceptive input data were three joint angles with added noise (shoulder1, shoulder2 and elbow - Fig. 4(a)). The visual input was an rgb camera mounted on the head of the robot, with 640×480 pixels definition. The tactile input was generated by multimodal skin cells distributed over the arm [32]. D. Learning g(x) from visual and proprioceptive data In order to learn the sensory forward model, we applied Gaussian process (GP) regression: gv(x) ∼ GP. We pro- grammed random trajectories in the joint space that resembled horizontal displacements of the arm. Figure 4(a) shows the extracted data: noisy joint angles and visual location of the end-effector, obtained by color segmentation. To learn the visual forward model sv = gv(x), each sample was defined as the input joint angles sensor values x = (x1, x2, x3) and the output sv = (i, j) pixel coordinates. As an example, Figure 4(b) shows the learned visual forward model by GP regression with 46 samples (red dots). It describes the horizontal mean and variance (in pixels) with respect to two joints angles. The GP learning and its partial derivative computation with respect to x is described in the Appendix B. (a) Data recorded example (joints angles + noise, end-effector visual, end-effector cartesian) and schematic picture of the 3-DOF. (b) Learnt gv(x) for visual horizontal location (c) Tactile (left) and visual (right) event trajectories Fig. 4. Collected data. (a) Joints angles, visual and ground truth information of the end-effector. (b) Mean and the variance computed by the GP for the visual horizontal location depending on two joints. (c) Touch patterns extracted from tactile (117 forearm skin cells) and visual sources. E. Extracting visuo-tactile data We used proximity sensing information (infrared sensors) from 117 different skin cells to discern when the arm was being touched. The sensor outputed a filtered signal ∈ (0, 1). From the other's hand visual trajectory and the skin proximity activation, we computed the level of synchrony between the two patterns (Fig. 4(c)). Timings for tactile stimuli st were obtained by setting a threshold over the proximity value: prox > 0.7 → activation. Timings for the other's trajectory events were obtained through the velocity components. Detected initial and end positions of the visual touching are depicted in Fig. 4(c) (right, green circles). V. RESULTS We compared the drifting data extracted from the rubber- hand illusion experiment in human participants and the robot. In order to obtain the robot results, we fixed in advance the model parameters for the learning and the body estimation stages. gv(x) ∼ GP learning hyperparameters: signal variance σn = exp(0.02) and kernel length scale l = exp(0.1). The integration step was ∆t = 0.05 (20Hz) and the error variances were σx ∈ R3 = [1, 1, 1], σp ∈ R3 = [1, 1, 1], σvt ∈ Z 2 = [2, 2]. The adaptability rate of µx was λ = 1. The visual feature from the real hand sv was not used in the rubber hand illusion experiment as the arm was covered. Finally, the visuo-tactile , where σt = function (Eq. 5) parameters were: b1 = σt d2 0.001 and dmax = 0.0016; b2 = 25; and a1 = a2 = 1. The robot drift was computed by subtracting the estimated end-effector position gv(x) and the ground truth location, and x was dynamically updated minimizing the prediction error using the proposed model. A. Comparative analysis max (a) Drift (b) Relative drift Fig. 5. Human proprioceptive drift vs end-effector robot estimation drift after the rubber hand illusion experiment. (a) Drift in cm. Positive values express displacement towards the fake hand. (b) Relative drift depending on the distance between fake and real arm. Figure 5 shows the proprioceptive drift comparison. Fig. 5(a) shows similar drifting patterns in both the robot and the human participants. A drift towards the fake hand emerges in both cases when the distance is small and then vanishes with longer distances. The prior information used for the tactile likelihood function parameters is modulated when the effect is taking place, as the error will start propagating when the gradient of the function is noticeable. Furthermore, the relative drift (Fig. 5(b)) also showed that, for close distances, the amount of displacement is the same, and then it decreases until vanishing. The robot was tested on even closer distances than humans, since the human experimental setup was not equipped for distances beneath 15cm. The large increase in proprioceptive drift for 10cm distance between fake and real hand is an interesting prediction for human data, that could be tested in future work. B. Human data analysis Data exceeding a range of two standard deviations around the mean was excluded from further analysis. T-tests were used to test the proprioceptive drift, the visual drift and the illusion score in each condition against zero. In the first three conditions the proprioceptive drift was significantly different from zero, while it was not in the other three conditions (Table I). Employing Bonferroni-Correction only leaves a trend towards significance in the 20cm distance condition. However, three conditions is still the average of the first significantly different from zero M : 13.72, SD : 14.84, p < .001) while the average of the other conditions is not (M 8.24 :, SD : 19.79, p > .05). The visual drift was only significantly different from zero in the 30cm distance condition (M : 14.01, SD : 29.72, p < .01). The illusion score was significantly different from zero in all conditions (all p < .05). Partial Pearson correlations between illusion score and proprioceptive drift, illusion score and visual drift and between proprioceptive drift and visual drift were not significant (all p > .05). top plot, also shows how smooth body configuration output µx1:3 is (blue line). The robot inferred the most plausible body joints angles given the sensory information, which in this case yielded a horizontal drift on the estimated end-effector location. A video of the evolution of the variables during the artificial rubber-hand illusion experiment can be found at http://web.ics.ei.tum.de/∼pablo/rubberICDL2018PL.mp4. DESCRIPTIVE AND INFERENCE STATISTICS FROM PROPRIOCEPTIVE DRIFT TABLE I DATA IN EACH CONDITION. Condition 15 cm 20 cm 25 cm 30 cm 35 cm 40 cm std mean df 14.89 mm 27.08 mm 19 12.79 mm 18.99 mm 17 13.22 mm 23.51 mm 16 15.52 mm 16 4.39 mm 18.00 mm 18 7.84 mm 5.45 mm 31.34 mm 17 t-value 2.46 2.86 2.32 1.17 1.90 0.74 p-value .024 .011 .034 .260 .074 .471 (a) End-effector drift depending on the sensing information (propio., visuo-tactile, and proprio+visuo-tactile) (a) propriceptive drift (b) illusion score Fig. 6. Mean and standard deviation of (a) the visual and the proprioceptive drift, and (b) the illusion score depending on the distance. C. Robot model analysis We analyzed the internal variables of the proposed model during the visuo-tactile stimulation and the induced end- effector estimation drift towards the fake arm. Figure 7(a) shows the robot camera view with the final end-effector estimation overlaid after 12 seconds. Depending on the dif- ferent enabled modalities (proprioceptive, visuo-tactile and proprioceptive+visuo-tactile), body estimation evolved differ- entially, accordingly the prediction of the end-effector gv(x). Fig. 7(b) shows the evolution of the body configuration in term of joint angles and the corresponding prediction errors. We did initialize the robot belief in a wrong body configuration to further show the adaptability of the model. During the first five seconds, the system converged to the real body configuration. Afterwards, when perturbing with synchronous visuo-tactile stimulation, a bias appeared on the body joints. This implies a drift of the robot end-effector towards the location of the fake arm visual feature. Tactile perturbations are shown as prediction error bumps (yellow line). Fig. 7(b), (b) Evolution of body estimation and predictive errors Fig. 7. Replicating the rubber-hand illusion on a robot. (a) End-effector estimation depending on the modality used. (b top) Joints angles in radians: real (black dotted line), estimated µx (blue dotted line) and current belief x (red line). (b middle) Errors between expected and observed values. (b bottom) g(x) values evolution during the experiment. In the first five seconds, in which there is no tactile stimulation, the estimation is refined. Next, we inject tactile stimulation while the experimenter is touching a fake arm. When visuo-tactile stimulation becomes synchronous, a horizontal drift appears on the end-effector estimation. VI. DISCUSSION: BODY ESTIMATION AS AN EXPLANATION FOR THE PERCEPTUAL DRIFT It has been shown that during the rubber hand illusion, the location of the real hand is perceived to be closer to the rubber hand than before. Similarly, the location of the rubber hand is mislocalized towards the real hand [7]. Our results from the robot and humans support the former finding: in our predictive coding scheme, the representation of both hands merged into a common location between both hands, due to inferring one's body's location from minimizing free energy. This body estimation generated a drift of the perceived location of the real hand towards the equilibrium location, which was visible in the data from both the humans and the robot. In comparative analysis, the patterns of the drift resembled each other, both in terms of absolute and relative values. The three closest tested distances showed a substantial proprioceptive drift. All other distances showed a smaller drift, approaching zero. For these, the distance between the fake and the real hand was probably too large for the fake hand to be fully embodied, supporting [6] simulation data exhibiting a reduced illusion probability for distances over 30cm. Although previous and the present research support predictive coding as a probable underlying mechanism of the rubber hand illusion, other accounts can not be ruled out by the present work. Human illusion score data, however, did not mirror the proprioceptive shift pattern found. For all distances, illusion scores ranged between 20 and 35, which on our continuous scale up to 100 resembles illusion scores previously found from 1 on a discrete scale to 3, e.g. [8]. Given this, we can assume that we were able to induce the illusion in every condition. Illusion scores and the proprioceptive drift, additionally, were not correlated. This supports the current debate that body-ownership illusions and the drift are two different, but related, processes [30]. The proprioceptive drift is an unconscious process - in contrast to the illusion, which is consciously accessible to the subjects. Hence, it might be possible that the predictive coding formulation in its uncon- scious form can explain drifting patterns, while it is not as such sufficient to explain body-ownership illusions. In contrast to [7], we did not find a conclusive visual drift in the human experiment. From participants' personal communication, we know that many used visual landmarks to estimate the position of the rubber hand, but of the real hand also. Differences in this strategy for localization would not only account for the high variance we observed in the drift data, but also for the small magnitude of the proprioceptive drift - as compared to the values reported in other studies (e.g. [16]). Beyond that, the method we used for localization is probably prone to high variance due to small mouse move- ments. Although we tried to account for that by repeating the localization ten times, more trials might be needed as performed in [6]. Arguably, however, the lack of visual drift in our study does not contradict the predictive coding scheme. Actually, some of our participants communicated that they experienced that the representation of both hands merged together. This is supported by the positive mean (14.16) of the actual control statement S10 "It felt as if the rubber hand and my own right hand lay closer to each other", which was even higher than the mean response (-4.95) to the actual illusion statement S3 "I felt as if the rubber hand were my hand". Further investigations, accounting for the variance in localization, are needed to support this conjecture. The computational model presented here also generates predictions about the temporal dynamics of the rubber hand illusion. The constant accumulation of information resulting in an also accumulated drift of the body estimation (see 7(a)) is comparable to findings from [9] (see 1), who also found an accumulation of the drift over time in humans. To provide a finer temporal comparison, the dynamics of the human illusion should be further investigated. VII. CONCLUSION We implemented the rubber hand illusion experiment on a multisensory robot. The perception of the real hand's position drifted towards the rubber hand, following a similar pattern in humans and the robot. We suggest that this proprioceptive drift resulted from a merging body estimation between both hands. This supports an account of the proprioceptive drift underlying body-ownership illusions in terms of the predictive coding scheme. Future work will address the mechanisms behind awareness of body-ownership. APPENDIX A FREE ENERGY GRADIENT p(xsp, sv) = p(spx)p(svx)p(svtx)p(x) (6) Applying logarithms we get the negative free energy formu- lation: F = ln p(spx) + ln p(svx) + ln p(svtx) + ln p(x) (7) Substituting the probability distributions by their functions f (.; .), and under the Laplace approximation [33] and assum- ing normally distributed noise, we can compute the negative free energy as: F = ln f (sp; gp(x), σp) + ln f (sp; gp(x), σp) + ln f (x; µx, σx) = − (x − µx)2 − (sp − gp(x))2 2σx + − (sv − gv(x))2 (cid:2)− ln σx − ln σsp − ln σsv − ln σsvt − (svt − gvt(x))2 (cid:3) + C. 2σvt 2σp 2σv (8) + 1 2 In order to find x in a gradient-descent scheme we minimize Eq. 8 through the following differential equation: x = − x − µx σx + sp − gp(x) + σp (cid:48) g p(x) + sv − gv(x) σv (cid:48) g v(x) + svt − gvt(x) σvt (cid:48) vt(x) g (9) In the case that x is equivalent to gp(x) like using the joint angles values directly as the body configuration, then the proprioceptive error can be rewritten as: sp−x and the gradient becomes 1. Generalizing for i sensors we finally have: si − gi(x) (cid:88) ∂gi(x)T = − (x − µx) (10) + σx ∂ x σi i ∂F ∂ x APPENDIX B GP REGRESSION Given sensor samples s from the robot in several body configurations x and the covariance function k(xi, xj), the training is performed by computing the covariance matrix K(X, X) on the collected data with noise σ2 n: ∀i, j ∈ x nI + k(xi, xj) kij = σ2 (11) The prediction of the sensory outcome s given x is then computed as [34]: g(x) = k(x, X)K(X, X)−1s = k(x, X)α (12) where α = choleski(K)T\(choleski(K)\s). Finally, in order to compute the gradient of the posterior g(x)(cid:48) we differentiate the kernel [35], and obtain its prediction analogously as Eq. 12: g(x)(cid:48) = ∂k(x, X) ∂ x K(X, X)−1s = ∂k(x, X) ∂ x α (13) Using the squared exponential kernel with the Mahalanobis distance covariance function, the derivative becomes: g(x)(cid:48) = −Λ−1(x − X)T (k(x, X)T · α) (14) where Λ is a matrix where the diagonal is populated with the length scale for each dimension (diag(1/l2)) and · is element- wise multiplication. REFERENCES [1] P. Lanillos, E. Dean-Leon, and G. Cheng, "Enactive self: a study of engineering perspectives to obtain the sensorimotor self through enaction," in IEEE Int. Conf. on Developmental Learning and Epigenetic Robotics, 2017. [2] M. Botvinick and J. Cohen, "Rubber hands feeltouch that eyes see," Nature, vol. 391, no. 6669, p. 756, 1998. [3] P. Lanillos, E. Dean-Leon, and G. Cheng, "Yielding self-perception in robots through sensorimotor contingencies," IEEE Trans. on Cognitive and Developmental Systems, no. 99, pp. 1–1, 2016. [4] K. Kilteni, A. Maselli, K. P. Kording, and M. Slater, "Over my fake body: body ownership illusions for studying the multisensory basis of own-body perception," Frontiers in human neuroscience, vol. 9, p. 141, 2015. [5] K. Friston, "A theory of cortical responses," Philosophical Transactions of the Royal Society of London B: Biological Sciences, vol. 360, no. 1456, pp. 815–836, 2005. [6] M. Samad, A. J. Chung, and L. Shams, "Perception of body ownership is driven by bayesian sensory inference," PloS one, vol. 10, no. 2, p. e0117178, 2015. [7] R. Erro, A. Marotta, M. Tinazzi, E. Frera, and M. Fiorio, "Judging the position of the artificial hand induces a visual drift towards the real one during the rubber hand illusion," Scientific reports, vol. 8, no. 1, p. 2531, 2018. [8] D. M. Lloyd, "Spatial limits on referred touch to an alien limb may reflect boundaries of visuo-tactile peripersonal space surrounding the hand," Brain and cognition, vol. 64, no. 1, pp. 104–109, 2007. [9] M. Tsakiris and P. Haggard, "The rubber hand illusion revisited: Visuotactile integration and self-attribution," Journal of experimental psychology. Human perception and performance, vol. 31, no. 1, pp. 80–91, 2005. [10] M. Botvinick and J. Cohen, "Rubber hands 'feel' touch that eyes see," Nature, vol. 391, no. 6669, p. 756, 1998. [11] K. Kilteni, A. Maselli, K. P. Kording, and M. Slater, "Over my fake body: Body ownership illusions for studying the multisensory basis of own-body perception," Frontiers in human neuroscience, vol. 9, 2015. [12] L. Aymerich-Franch, D. Petit, G. Ganesh, and A. Kheddar, "Non- human looking robot arms induce illusion of embodiment," International Journal of Social Robotics, vol. 9, no. 4, pp. 479–490, 2017. [13] M. Tsakiris, "My body in the brain: A neurocognitive model of body- ownership," Neuropsychologia, vol. 48, no. 3, pp. 703–712, 2010. [14] H. H. Ehrsson, C. Spence, and R. E. Passingham, "That's my hand! activity in premotor cortex reflects feeling of ownership of a limb," Science, vol. 305, no. 5685, pp. 875–877, 2004. [15] A. Kalckert and H. H. Ehrsson, "The onset time of the ownership sensation in the moving rubber hand illusion," Frontiers in psychology, vol. 8, p. 344, 2017. [16] R. Zopf, G. Savage, and M. A. Williams, "Crossmodal congruency measures of lateral distance effects on the rubber hand illusion," Neu- ropsychologia, vol. 48, no. 3, pp. 713–725, 2010. [17] C. Preston, "The role of distance from the body and distance from the real hand in ownership and disownership during the rubber hand illusion," Acta psychologica, vol. 142, no. 2, pp. 177–183, 2013. [18] S. C. Pritchard, R. Zopf, V. Polito, D. M. Kaplan, and M. A. Williams, "Non-hierarchical influence of visual form, touch, and position cues on embodiment, agency, and presence in virtual reality," Frontiers in psychology, vol. 7, p. 1649, 2016. [19] N. Ratcliffe and R. Newport, "The effect of visual, spatial and temporal manipulations on embodiment and action," Frontiers in human neuro- science, vol. 11, p. 227, 2017. [20] T. R. Makin, N. P. Holmes, and H. H. Ehrsson, "On the other hand: Dummy hands and peripersonal space," Behavioural brain research, vol. 191, no. 1, pp. 1–10, 2008. [21] M. S. Graziano, D. F. Cooke, and C. S. Taylor, "Coding the location of the arm by sight," Science, vol. 290, no. 5497, pp. 1782–1786, 2000. [22] H. H. Ehrsson, N. P. Holmes, and R. E. Passingham, "Touching a rubber hand: Feeling of body ownership is associated with activity in multisensory brain areas," Journal of neuroscience, vol. 25, no. 45, pp. 10 564–10 573, 2005. [23] A. Guterstam, G. Gentile, and H. H. Ehrsson, "The invisible hand illusion: Multisensory integration leads to the embodiment of a discrete volume of empty space," Journal of cognitive neuroscience, vol. 25, no. 7, pp. 1078–1099, 2013. [24] A. Guterstam, H. Zeberg, V. M. Ozc¸iftci, and H. H. Ehrsson, "The mag- netic touch illusion: A perceptual correlate of visuo-tactile integration in peripersonal space," Cognition, vol. 155, pp. 44–56, 2016. [25] M. A. J. Apps and M. Tsakiris, "The free-energy self: A predictive coding account of self-recognition," Neuroscience and biobehavioral reviews, vol. 41, pp. 85–97, 2014. [26] D. Zeller, K. J. Friston, and J. Classen, "Dynamic causal modeling of touch-evoked potentials in the rubber hand illusion," NeuroImage, vol. 138, pp. 266–273, 2016. [27] P. Lanillos and G. Cheng, "Adaptive robot body learning and estimation through predictive coding," arXiv preprint arXiv:1805.03104, 2018. [28] R. Bogacz, "A tutorial on the free-energy framework for modelling perception and learning," Journal of mathematical psychology, 2015. [29] M. P. M. Kammers, F. de Vignemont, L. Verhagen, and H. C. Dijkerman, "The rubber hand illusion in action," Neuropsychologia, vol. 47, no. 1, pp. 204–211, 2009. [30] Z. Abdulkarim and H. H. Ehrsson, "No causal link between changes in hand position sense and feeling of limb ownership in the rubber hand illusion," Attention, perception & psychophysics, vol. 78, no. 2, pp. 707– 720, 2016. [31] E. Dean-Leon, B. Pierce, F. Bergner, P. Mittendorfer, K. Ramirez- Amaro, W. Burger, and G. Cheng, "Tomm: Tactile omnidirectional mobile manipulator," in Robotics and Automation (ICRA), IEEE Int. Conf. on, 2017, pp. 2441–2447. [32] P. Mittendorfer and G. Cheng, "Humanoid multimodal tactile-sensing modules," IEEE Trans. on robotics, vol. 27, no. 3, pp. 401–410, 2011. [33] K. Friston, "Hierarchical models in the brain," PLoS computational biology, vol. 4, no. 11, p. e1000211, 2008. [34] C. E. Rasmussen and C. K. I. Williams, Gaussian Processes for Machine The MIT Learning (Adaptive Computation and Machine Learning). Press, 2005. [35] A. McHutchon, "Differentiating gaussian processes," 2013.
1803.05840
1
1803
2018-03-15T16:25:47
Effective Connectivity from Single Trial fMRI Data by Sampling Biologically Plausible Models
[ "q-bio.NC", "physics.data-an" ]
The estimation of causal network architectures in the brain is fundamental for understanding cognitive information processes. However, access to the dynamic processes underlying cognition is limited to indirect measurements of the hidden neuronal activity, for instance through fMRI data. Thus, estimating the network structure of the underlying process is challenging. In this article, we embed an adaptive importance sampler called Adaptive Path Integral Smoother (APIS) into the Expectation-Maximization algorithm to obtain point estimates of causal connectivity. We demonstrate on synthetic data that this procedure finds not only the correct network structure but also the direction of effective connections from random initializations of the connectivity matrix. In addition--motivated by contradictory claims in the literature--we examine the effect of the neuronal timescale on the sensitivity of the BOLD signal to changes in the connectivity and on the maximum likelihood solutions of the connectivity. We conclude with two warnings: First, the connectivity estimates under the assumption of slow dynamics can be extremely biased if the data was generated by fast neuronal processes. Second, the faster the time scale, the less sensitive the BOLD signal is to changes in the incoming connections to a node. Hence, connectivity estimation using realistic neural dynamics timescale requires extremely high-quality data and seems infeasible in many practical data sets.
q-bio.NC
q-bio
Effective Connectivity from Single Trial fMRI Data by Sampling Biologically Plausible Models H.C. Ruiz Euler, H.J. Kappen August 19, 2020 Abstract The estimation of causal network architectures in the brain is fundamental for understanding cognitive information processes. However, access to the dynamic processes underlying cognition is limited to indirect measurements of the hidden neuronal activity, for instance through fMRI data. Thus, estimating the network structure of the underlying process is challenging. In this article, we embed an adaptive importance sampler called Adaptive Path Integral Smoother (APIS) into the Expectation-Maximization algorithm to obtain point estimates of causal connectivity. We demonstrate on synthetic data that this procedure finds not only the correct network structure, but also the direction of effective connections from random initializations of the connectivity matrix. In addition -- motivated by contradictory claims in the literature -- we examine the effect of the neuronal time scale on the sensitivity of the BOLD signal to changes in the connectivity and on the maximum likelihood solutions of the connectivity. We conclude with two warnings: First, the connectivity estimates under the assumption of slow dynamics can be extremely biased if the data was generated by fast neuronal processes. Second, the faster the time scale, the less sensitive the BOLD signal is to changes in the incoming connections to a node. Hence, connectivity estimation using realistic neural dynamics time scale requires extremely high quality data and seems infeasible in many practical data sets. Introduction 1 In recent years, the field of neuroimaging has seen a rapidly increasing interest in effective connectivity estimations from functional magnetic resonance imaging (fMRI) data. Although this data acquisition method is very powerful to investigate human brain function by identifying brain regions that are active during perceptual or cognitive tasks, fMRI time series are an indirect, delayed and blurred measurement of the actual signal of interest, the latent neuronal activation of a specific region of interest (ROI). Thus, estimating the underlying connectivity is challenging. Roughly speaking there are two approaches to this problem based on the distinction between functional and effective connectivity. The first approach seeks to estimate the temporal correlations between separable ROIs [7]. Examples of these, which are known to correctly estimate the network structure, are partial correlation, regularized inverse covariance and some Bayes net methods [28]. These symmetric measures, however, have no information about the directionality of the connection, i.e. the influence that one node exerts over another. This "effective" connectivity is in general harder to estimate but it is often of great interest. We can distinguish between two families of approaches; purely data-driven methods that attempt to infer directionality directly from the time series using statistical measures and methods based on dynamic models that seek to find a forward model to fit the data. Purely data-driven methods fall generally into three classes. First, lag-based methods, e.g. Granger causality [13, 12, 5], are variations of the well-known auto-regressive models. In this framework one considers the similarity between a pair of time series, one of which is shifted in time. If the lagged time series helps predict the zero-lagged time series, then a causal relation is inferred. A second family of methods is based on the concept of conditional independence and follow the ideas from structural learning in Bayes networks [14, 22]. The third class considers higher order statistics to infer causality. For instance, Patel's τ approach measures the asymmetry between the conditional probabilities of functionally connected nodes. If the activation probability of a node 1 given node 2 is larger than the activation probability of node 2 given node 1, this is interpreted as a directed connection 2 → 1 [20]. Although many of these methods are widely used, the estimation of directed connectivity proves difficult with these methods [28]. Dynamic model methods frame the inference problem as a state-space model (SSM). The latent states are the dynamic degrees of freedom of the process underlying the data. Hence, in general, estimating the parameters of the 1 model requires also the estimation of the hidden process, which is in most cases analytically infeasible and some approximations are needed. Once the latent process estimation is addressed, one typically uses the Expectation- Maximization (EM) algorithm in either its Maximum-Likelihood (ML) or variational Bayes versions [26, 25]. For fMRI data, model-based methods usually distinguish themselves in the approximations they make of the dynamic system or the latent process estimation (E-step). For instance, in [25] the fMRI time series is modeled as a linear convolution of the bi-linear latent system with the canonical hemodynamic response function (HRF) and its time derivative [21]. This approximation allows using the Kalman smoother for the E-step. On the contrary, dynamic causal modeling (DCM) uses a biologically plausible nonlinear model for the hidden process. Here, the neuronal activity is given as a bi-linear system coupled to the Balloon model describing the nonlinear relation between blood flow and volume [2, 18, 10]. In addition, the fMRI observations are modeled as a nonlinear function of the hidden variables. Hence, to deal with the nonlinearities of the biological model, DCM resorts to either a Volterra expansion of the dynamic system or the generalized filtering method [8, 6, 9]. In addition, DCM uses a Gaussian approximation of the posterior over the parameters and its mean and covariance are estimated using EM update rules. For this, the adjacency matrix is specified beforehand, effectively imposing a strong prior on the connectivity of the system. Using the log-evidence, the proposed models are scored to find the best model amongst them [9]. Although DCM has been widely used, concerns about this approach have spurred a discussion about its feasibility and validity due to the combinatorial explosion of the network structure and the apparent inability of the scoring procedure to distinguish between generally accepted networks and randomly generated ones [17, 1, 4]. In this article, we present an alternative method for connectivity estimation. Our approach is similar to DCM in the sense that it uses the same model but differs in two important ways. First of all, in the E-step the full posterior over the latent process is estimated using an optimal control approach that was first introduced in [23]. Second, in the M-step the connectivity is optimized without prior assumptions as in [9, 8]. The Monte Carlo estimates involved in these steps are the only approximations required and it is proven to work for nonlinear deconvolution of fMRI time series [24]. We show using synthetic data that the connectivity estimates obtained are close to the ground truth and that adding a small L1-regularization on the connections is beneficial to obtain sparse estimates that generalize better on unseen time series. Furthermore, the proposed method obtains estimates from single event fMRI time series that are robust against random initializations of the connectivity matrix. In addition, we study the sensitivity of the BOLD signal depending on the neuronal time scale and the effect of this parameter on the connectivity estimates. This analysis is motivated by the different claims about the value of the neuronal time scale in the literature. For instance, [28] used a time constant resulting in a mean neural lag of 50 ms and it is argued that this value is towards the upper limit of observed lags in general. Contrary to this claim is the assumption made in the DCM literature, where it is argued that the scale ought to be of order 1 s [8, 9]. We observe a significant effect of the neuronal time scale on the connectivity estimates and a remarkable lower sensitivity of the BOLD signal to changes in the connectivity for fast neuronal dynamics. Hence if the underlying neuronal processes are fast, the quality of the data must be significantly higher to obtain reliable, unbiased estimates of the effective connectivity. 2 Method 2.1 Modeling fMRI Data Similar to DCMs [8], we consider the fMRI forward model as a network of m regions z = (z1, z2, . . . , zm) that follow stochastic dynamics given by dzt = A(Czt + u(zt, t) + BIt)dt + (1) where dWt ∼ N (0, dt) is a m-dimensional Wiener process with variance1 dt = 0.01 and σz > 0. The parameter A sets the time scale of the neuronal response, while the connectivity matrix C has diagonal with −1 and off-diagonal elements Cij ≤ 1, i (cid:54)= j. This form assumes that the fastest scale is the within-node temporal decay. The term √ A in the diffusion ensures that the stationary distribution remains invariant under changes in the time scale A. Contrary to DCMs, the input strength has been rescaled such that the stationary point of the system with non-zero constant input is independent of A. Hence, the inverse time scale A determines only how fast the neuronal system follows the input. AσzdWt √ 1In all our simulations we used this discretization step unless stated otherwise. 2 Parameter Value Parameter A  τs τf τ0 1 Hz 0.8 1.54 2.44 1.02 σz E0 k1 k2 k3 Value 10−3 0.4 7E0 2 2E0 − 0.2 Parameter Value Parameter Value √ 0 σz/ 2 TBD TBD 0.002 2.5 0.32 0.018 B α−1 V0 µs,0 µz,0 σz,0 σs,0 σf,q,v,0 µf,q,v,0 0 1 σy Table 1: Parameters for the neural dynamics (top row) and for the BOLD transformation (bottom rows). Notice that we define the process in equation (1) to have an unknown function u(z, t). This function is the importance sampling controller learned with APIS. In general, u(z, t) can be any parametrized function [?], but in this paper it is chosen to have the simple form u(z, t) = az + bt with a a constant and bt a time varying function. We assume that the external input It and its strength B are known. Each node's activity zi is coupled to a nonlinear deterministic system modeling the hemodynamic transformation. For each node i there are two Hemodynamic equations [10] (cid:19) dt (cid:18) zi − si τs − fi − 1 τf dsi = dfi = sidt (2) (3) and two equations of the Balloon model2 [2], dqi = dvi = 1 τ0 1 τ0 The BOLD signal change is given by (cid:18) 1 − (1 − E0)1/fi fi (fi − vα i ) dt. E0 (cid:19) dt − vα−1 i qi v (cid:0)1 − q (cid:2)k1 (1 − qt) + k2 yi(t) = B(qi, viθ) + σydWy (cid:1) + k3 (1 − v)(cid:3), dWy ∼ N (0, 1) and θ denotes all parameters of the where B(qt, vtθ) := V0 system. For simplicity, we assume the same hemodynamic transformation for all nodes since our focus is on connectivity estimates. The parameters are found in [10]. A summary of all the values used in this article is given in table 1, unless a different value is stated explicitly. For now, we restrict our attention to slow neuronal processes with a neuronal lag of A = 1 Hz, which is around the typical values in the literature on effective connectivity, e.g. [8, 9]. The input strength B is chosen such that the resulting amplitude of BOLD responses are around 2-4%. In addition, we consider for the prior over the initial states in each node xi,0 = (zi,0, si,0, fi,0, qi,0, vi,0) a normal distribution3 with mean µ0 = (µz,0, µs,0, µf,0, µq,0, µv,0) and a covariance given by a diagonal matrix with entries (z,s,f,q,v),0 set to be the variance of the stationary distribution induced by the Ornstein-Uhlenbeck process in σ2 √ equation (1) when It = u(z, t) = 0. Hence, the variance for p(z0) is set to σz,0 = σz/ 2. Since all hemodynamic variables are deterministic, their variance is estimated by forward sampling. (4) Now that the model has been described, we proceed with a more detailed explanation of the parameter inference procedure given by the EM-algorithm. 2.2 EM-algorithm for Time Series The Expectation-Maximization algorithm [3] is an iterative method to find maximum-likelihood or maximum a posteriori (MAP) estimates. There are several applications to state space models (SSM), e.g. [27]. In its general form, its objective is to maximize a lower bound of the marginal likelihood p(y0:Tθ) = dz[0:T ]p(z[0:T ], y0:Tθ) where y0:T = {y(tk)k = 0, . . . , K} is the time series, θ the set of parameters to estimate and z[0:T ] = {ztt ∈ [0, T ]} are ´ 2For clarity in the notation, we denote the exponent of the volume fraction 1/α simply as α. 3Notice that due to the small discretization step dt and noise levels used here, the log-transformation of the hemodynamic variables was not required [29]. Nevertheless, it is straightforward to use this transformation in our procedure. 3 the hidden (continuous) processes underlying the observations. For any probability density q(z[0:T ]) we get from Jensen's inequality (cid:2)q(z[0:T ])p(z[0:T ], y0:Tθ)(cid:3) is the Kullback-Leibler (KL) divergence. Hence, the variational density where DKL q(z[0,T ]) must be optimized to tighten this inequality. It turns out that the optimal variational density q(z[0:T ]) for a fixed value of the parameters θ = θ is precisely the posterior p(z[0:T ]y0:T , θ) [19]. Hence, given an initial value θ, the objective is (cid:2)q(z[0:T ])p(z[0:T ], y0:Tθ)(cid:3) log [p(y0:Tθ)] ≥ −DKL (cid:105)(cid:105) (cid:104)−DKL (cid:104) θ∗ = argmax θ p(z[0:T ]y0:T , θ)p(z[0:T ], y0:Tθ) (5) This algorithm entails two steps in each iteration. First, the E-step is used to obtain expectations with respect to the posterior over the latent process z[0,T ]. Then, the M-step is a gradient update to maximize the objective function with respect to the parameters in question. 2.2.1 E-step: APIS Given the above model, we can sample from the posterior using APIS. This adaptive importance sampling method samples N trajectories, or particles, by forward integration of (1)-(3) initialized with u(z, t) = 0. Then, a total cost Sξ is assigned to each hidden process ξ [11, 15, 16] via (cid:104) 0 Sξ = − T dt log g(y0:Tz(ξ) [0,T ]) (cid:105) − 1 2 t )2 − (u(ξ) A σ2 z √ A σz u(ξ) t dW (ξ) t , t) and g(y0:Tz(ξ) t t denotes the noise realization of process ξ, u(ξ) where dW (ξ) [0,T ]) is the likelihood with respect to the latent process z(ξ) [0,T ] given by equation (4). We call the first term of Sξ the "state cost" of particle ξ. The negative of the total cost exponentiated gives the unnormalized importance weight used in the Monte Carlo estimates to learn the control function u(z, t). After u(z, t) is updated, it is used to sample again forward particles. These steps are repeated until the effective sample size (ESS) converges or reaches a predefined value. For details on APIS we refer the reader to [23]. t = u(z(ξ) APIS not only increases the efficiency of sampling, but it also minimizes the mean cost S of the particle ensemble. This ensures that the posterior neuronal and BOLD signals give the best possible fit, even if there is a strong mismatch between the model and the ground truth [24]. In addition, the control appears in the computation of the gradients of the KL divergence, giving reliable Monte Carlo estimates in the M-step. For a 3 node network, we use N = 20000 particles over 500 iterations to ensure that the scheme has sufficient time to bootstrap the E-step and converge to the MAP connections in the M-step. Depending on the initialization, the estimations can take more or less iterations, but we find that this number allows for convergence in most cases. A learning rate of ηE = 0.1 achieves a fast bootstrapping while maintaining the estimations of the control signals stable. For the annealing threshold we use γE = 0.02 − 0.03, corresponding to 2-3% of effective samples used for the estimations during bootstrapping. 2.2.2 M-Step: Gradient Ascent in the Connectivity In the M-step, we maximize equation (5) with respect to the connectivity matrix C and assume that all other parameters have correct values. Since C is assumed to be time independent, this optimization requires estimates of zt at all time. Thus, we wish to maximize at the n-th iteration, E(cid:110) log [p(ztzt−dt, C)]y0:T , C (n)(cid:111) T(cid:88) t=dt Qn(C) = where the expectation is over the posterior latent process. Using equation (1) and the conditional independence of the components zi t given zt−dt, the above results in T(cid:88) m(cid:88) t=dt i=1 E (cid:26) −1 2σ2 z dt (cid:0)dzi t − Fi(zt, C)dt(cid:1)2 y0:T , C (n) (cid:27) Qn(C) = 4 where F (zt, C) = A(Czt + BIt). With ∂ikF := ∂F/∂Cik = Azk finally obtain T(cid:88) t=dt E (cid:26) A (cid:16) σ2 z ∂ikQn = Aui tdt + AσzdW i t t and dzt − F (zt, C)dt = Autdt + (cid:17) (cid:27) t y0:T , C (n) zk . √ √ AσzdWt we This gradient is used to update the connectivity at each iteration, C (n+1) ik = C (n) ik + ηM ∂ikQn where ηM is the learning rate. Note that both the E- and the M-step involve optimization. In the E-step we optimize the importance sampler for given C (n). In the M-step, we optimize C (n). In practice, we find that it is beneficial to wait until the ESS in the E-step is sufficiently high. This is ensured by imposing a threshold γM on the ESS above which the connectivity is updated. Empirically, it is found that γM ∈ [2γE, 5γE] and a learning rate of ηM = 0.01 work well. The latter is usually found such that after each update, the ESS takes only few iterations to reach γM again. Naturally, there is a trade-off between large updates of the connectivity and the stability of the ESS. In addition, we use momentum with a rate of κ = 0.9 to improve the gradient ascent procedure. Remember that the within-node time scale is assumed to be the fastest time scale in the network and all nodes have the same value. Hence, we fix the diagonal elements of C to −1. Although they are not optimized, the above procedure can be readily extended to estimate these elements as well. The initial random matrices for n = 0 are generated with the following procedure similar to [17]. First, on-edges are sampled with probability p1 = 0.5. This gives a random graph that defines the adjacency matrix, i.e. the directed connections among the ROIs. Then, the strength of these connections are randomly chosen from a uniform distribution on a small interval [c, d] where c , d ≤ 1. Here, the bound on the strength are chosen to be c = −0.5 and d = 0.5. This ensures that the sampled matrix has negative eigenvalues with very high probability, which is important to guarantee the stability of the latent process. The resulting matrix is accepted if all eigenvalues are negative, otherwise, the procedure is repeated. 2.3 Studying the Effects of the Neuronal Time Scale on Effective Connectivity Esti- mates The use of slow dynamics for the neuronal processes underlying fMRI data is argued in [9], but it stands in contrast to the case studied in [28] with a neural lag of approximately 50 ms, corresponding to A = 20 Hz. This is a significant difference in the assumptions made on the generative process of the data, hence, it is important to understand the consequences of these assumptions for effective connectivity estimates. For this, we study two important effects that depend on the time scale of the neuronal activity, the sensitivity of the BOLD response to changes in the connectivity and the maximum-likelihood solutions under different assumptions of the time scale. As a measure of sensitivity, the difference in the mean and variance of the posterior BOLD between two significant different models is computed. We call this difference the "sensitivity" of the mean sµ and variance sσ respectively. Hence, if the change in the response is large when moving from one connectivity model to another, we say that the system is highly sensible to changes in the connectivity. On the contrary, small changes in the BOLD response will make the discrimination between connectivity models more difficult. In addition, the inverse time scale A not only affects the sensitivity of the BOLD response, but also its overall shape and delay. This could have significant effects on the connectivity estimates. Thus, we study the solutions depending on the assumed neuronal time scale given that the underlying generative process has fast neuronal activation, i.e. we consider a model mismatch between the ground truth inverse time scale AGT and the assumed one for the reconstruction of the network A. For simplicity, we consider the deterministic system because we find that the variance of the posterior contributes marginally to the neg. log-likelihood (sσ (cid:28) sµ) and the posterior mean follows the same dynamics as the determin- istic system for sufficiently small noise levels. Hence, since the only randomness is in the observations, we do not use the EM approach described above. Instead, we use a brute force search on the neg. log-likelihood landscape. This approach requires the discretization of the connectivity space and, for each grid point, the integration of the full system on the time interval [0, T ]. In this analysis, consider for the ground truth a chain network I → 1 → 2 → 3 with connectivity matrix CGT,c and input I. To lower the computational effort in finding the connections we restrict the problem by fixing all connections but two, say (C31, C32). The brute force approach even in this restricted scenario is computationally too expensive to study the grid search solutions for a large number of time series. Hence, the search on this plane 5 is constrained further to the line C31 = −gC32 + h on which, given the input, the fix point of the neuronal system is invariant to changes in the connectivity. This constrain gives g = (CGT,c)23 and h = g · (CGT,c)32. To see this, consider the fix point solution z∗ of equation (1) with u(z, t) = 0, σz = 0. For node 3, the dependence of z∗ 3 on the connections (C31, C32) is given by z∗ 1 is the fix point of the activity in node 1. Keeping the input fixed sets the value of z∗ 3 will be completely determined by the proportionality factor h = gC32 + C31. By setting h = g · (CGT,c)32, we ensure that the fix point of the system remains constant at the value of the ground truth while we vary the connections on this line, which we call the z∗-line. Hence, the only relevant information to differentiate between connections is in the transients given by the values on this line. We will denote estimates obtained by the brute force search on the z∗-line as grid search connections to distinguish them from the EM approach described above. For a validation on this procedure we refer the reader to the appendix. 3 = [gC32 + C31]z∗ 1, so z∗ 1, where z∗ We use the grid search on two studies on the effect of the neuronal time scale on the grid search connections and the measurement precision needed to obtain reliable connectivity estimates. In these studies, we perform a brute force search on the z∗-line as follows. Given a value C32, we integrate the system to obtain the BOLD response, which is used to compute the log-likelihood and find the grid search solution of 1000 noise realizations, first keeping σy fixed and varying A, then keeping A fixed to the correct value AGT and varying σy. Notice that the variance observed in this analysis is clearly an under estimate of the true variance of the grid search solutions in the full connectivity space, because we restricted ourselves to a single line C31 = −gC32 + h in this space. However, the results of this analysis capture already the consequences of fast neuronal dynamics for connectivity estimates. 3 Results 3.1 Single Trial Estimates of Effective Connectivity from Slow Neuronal Activation In the following analysis we consider two connectivity structures with different topologies with no bidirectional connection, a chain network CGT,c and a triangle network CGT,∆. In both cases, we consider an external input to node 1 given by a box-car function It,1 = Θ[t − ton]Θ[tof f − t] where (ton, tof f ) = (3.2, 4.70) seconds and Θ is the Heaviside function. To generate the data from both examples CGT,c and CGT,∆, the systems are integrated forward T = 16 seconds with the initial state set to the mean value µ0. In each case, the response signals of the nodes are subsampled with a T R = 0.4 seconds, resulting in 3 time series each with 41 observations corrupted by Gaussian noise (σy = 0.002). This noise level is chosen to be 5-8% of the BOLD change, roughly the same as in [9]. are assumed, we use an L1-regularization term Σi,jλL1 Cij in equation (5). We illustrate how the EM-algorithm together with APIS obtains the correct directed edges. Since sparse networks For the first example, we consider the chain network with connections  −1 CGT,c = 0 0 0.5 −1 0 0.75 −1 0  . In figure 1, the connectivity estimation for the chain network C c λL1=0.05 is shown over iterations. In this case, we choose a regularization of λL1 = 0.05. Each graph in the panels corresponds to a different random initialization of the connectivity matrix. At the beginning, the connections do not change because the sampler must learn an appropriate controller to bootstrap the procedure. Once the controller is estimated with sufficient accuracy, the ESS surpasses a threshold and the connectivity is updated via gradient ascent, hence the sudden jumps in the connections. After a few jumps, the learning procedure stabilizes and the connections are updated at each iteration. In most cases, it takes around 200-250 iterations to converge to the MAP solution. used. The mean state cost of the sampled process E(cid:0)log(cid:2)g(y0:Tz[0,T ])(cid:3)(cid:1) decreased in all cases from around 1000 The ESS after convergence of most examples increased from less than 1% to higher than 90% of total particles to 68, meaning that the model found fits the data well. Notice that the randomly initialized matrices converged all to the same values close to the ground truth. The 31 is sufficiently weak and could be regarded as non existent since the other estimated connections connection C c 32 can be understood from the L1- dominate by an order of magnitude. The underestimation of the connection C c penalty in the M-step to obtain sparse solutions. This regularization shrinks the connections by penalizing their absolute value. In turn, this shrinkage is compensated by an increase in the other incoming connection to node 3 to reduce the state cost, i.e. C c 31 becomes non-zero. This hints to a strong regularization. To see the effect of strong regularization, we study the dependency of the solutions on the regularization strength λL1. 6 Figure 1: MAP-solutions for chain network: The figure shows the consistency of the estimation procedure against random initializations of the connectivity matrix. Each line represents the estimation procedure over iterations of a randomly initialized example. The regularization strength is set to λL1 = 0.05. All runs converged to the same matrix C c λL1=0.05. The estimated connections are not identical to the ground truth because of the finite data used in reconstruction. However, the qualitative structure of the network, including which connections are present and their directionality, is correctly reconstructed. On the lower right panel only an example of the upper off-diagonal connections is shown. In all cases they are estimated to vanish with very high precision. 7 Figure 2: MAP-solutions for chain network vs L1-regularization strength: a small amount of regularization λL1 = 0.01 improves the estimation of the connectivity compared to the unregularized case. The mean validation error of the BOLD signal over 20 new time series decreases compared to the unregularized case but increases rapidly with stronger regularization (right). Taking the connectivity at the minimum of the validation error gives us a connectivity with a small error relative to the ground truth connection (black dashed line, left panel) For a fixed initialization of uncoupled nodes C c 0 = −13×3, the connections of the chain network are estimated using different values λL1 ∈ [0, 0.1]. The analysis shows -- left panel of figure 2 -- that the ground truth connections (black dashed line) are consistently revealed across a wide range of λL1-values. As expected, zero regularization gives small non-zero contributions in edges that would not be allowed, but still the ground truth connections are easy to differentiate. Hence, although the L1-regularization is not crucial in this example, we see that applying a small regularization helps shrinking the value of non existent connections towards zero, which improves the estimate of the network making it generalize better, as seen on the right panel of figure 2, where the mean validation error is shown. To estimate the validation error we generate 1000 new time series using the ground truth connection and computed the mean negative log-likelihood or sate cost of the response obtained from each connectivity. This is a measure of how well our estimates fit the new data. We observe a decrease for weak regularization λL1 = 0.01 vis- ï¿oe-vis the unregularized case and a strong increase in the validation error for increasing λL1. From the minimum of this profile, we conclude that a weak regularization is beneficial, in this case using λL1 = 0.01 we obtain a solution pinpointing the ground truth to Finally, we proceed with the estimation of the triangle network to show that we obtain again the ground truth structure with a small regularization of λL1 = 0.01. In figure 3, we observe a robust estimation against random initializations of the connectivity matrix and a clear distinction between correct and incorrect edges and directions. In this case, all estimated connections are around C c λL1=0.01 = 0 0 0.45 −1 0 0.78 −1 0 CGT,∆ = 0 0 0.5 −1 0 0.5 −1 0.375 C ∆ λL1=0.01 = 0 0 0.48 −1 0.03 0.4 −1 0.41  .   .  −1  −1  −1 8 Figure 3: Triangle network with similar connection strengths as the chain network: The MAP-solution is again robust against random initializations of the connectivity and we can clearly distinguish between existent and non- existent edges and directions. Interestingly, in this case the overestimated connection is C ∆ 23 with the same order of magnitude as C ∆ 31 in the previous example. Again, we may vary λL1 to select its value by cross validation. Identifying Networks with Fast Neuronal Activations 3.2 3.2.1 BOLD Sensitivity to Changes in the Connectivity Diminishes with Fast Neuronal Dynamics The connectivity estimation above assumed that the temporal scale of the neuronal dynamics is of the order of a second. Here, we study the sensitivity of the BOLD response to changes in the connectivity as a function of A. We consider 3 different time scales A = 1, 5 and 50. The neuronal noise level σz = 0.01 is chosen such that the prior process has a standard deviation of 0.7%. The learning rate of APIS has to be adapted for the different values of A, roughly with the relation ηE = 0.1/A. All other parameters of APIS are kept fixed. Given the data generated by the chain network CGT,c, the posterior statistics of the BOLD signal are estimated in two extreme cases using APIS; one case defines the structure 3 ← 1 → 2 and the other is close to the ground truth CGT,c. The connectivity matrices are chosen such that the only difference to the ground truth are the connections into node 3,  .  −1 C1 = 0 0 0.5 −1 0 0 −1 0.375  ; C2 =  −1 0 0 0.5 −1 0 0.63 −1 0.06 In figure 4, we show the sensitivity of the BOLD response in node 3 when changing the connections from C1 to C2. We appreciate two characteristics of the posterior BOLD response. First, on the right panel, it is apparent that the variance of the signal has little sensitivity to the connectivity, with a change several orders of magnitude smaller than the change of the mean signal on the left. This suggests 9 Figure 4: Sensitivity of the BOLD posterior to changes in the connectivity. Notice that the change in the variance of the posterior BOLD (right) is five orders of magnitude smaller than the change in the posterior mean (left). Hence, faster neuronal activity implies less sensitivity of the BOLD signal to changes in the connectivity. The dashed black line is the noise level σy = 0.002 of the data generated with C c GT . that -- for small neuronal noise -- the contribution of the variance to the identification of the connectivity is marginal and we can focus on the mean signal, which is equal to the BOLD signal of a deterministic system. Second, on the left panel, for slow dynamics the amplitude of the mean signal change is larger than the observation noise level σy. However, already for A = 5 the sensitivity decreases below the observation noise and it becomes an order of magnitude smaller for large A. Hence, the faster the neuronal dynamics, the less sensible the BOLD response becomes to changes in the connectivity. The lack of sensitivity for faster time scales implies flat likelihood functions that contain no information to differentiate connectivity structures. Thus, A should be sufficiently small such that the sensitivity is higher than the observation noise. This seems to be a fundamental problem to reconstruct connectivity. One solution to this problem is more data, which effectively reduces the measurement noise. Another, option is to reconstruct with small A and hope for the best. Thus, we study the grid search solutions obtained for different inverse time scales A by the brute force search described before. 3.2.2 Connectivity Estimates Depend on the Assumed Neuronal Time Scale We study now the connectivity estimates as a function of the inverse neuronal time scale A. For each value of A, we find the connections from 1000 different time series with the ground truth given by the chain network CGT,c with the same parameters as before but AGT = 35. Figure 5 shows the mean and standard deviation of these estimates. Taking the amplitude of sµ in figure 4 as a rough upper limit for the noise level that allows connectivity estimations, the analysis is performed with two noise levels, σy = 2 × 10−4 (on the top panels) and σy = 2 × 10−3 (on the bottom panels). Upon examination of the top panels, we reinforce our expectation that with sufficient precision, the variance of the grid search solutions is small. Interestingly, the inverse time scale A biases the connectivity estimates in a significant way. Nevertheless, it is possible to obtain in principle the correct connectivity if the time scale is jointly estimated because there is a clear minimum in the negative log-likelihood profile (red dot on the right panel). On the contrary, with a noise level an order of magnitude higher, the identification of connections is problematic. Although the bias of the estimations is the same as above, the variance of the grid search solutions increases much faster with the inverse time scale because of the flat profiles of the neg. log-likelihood function. Thus, the grid search solutions spread across a wide range of values, making it difficult to pinpoint the connectivity. As a consequence, for A ≥ 15 Hz identifying the correct connections will not be possible, even after learning the inverse time scale. This can be appreciated by the flat profile of the mean neg. log-likelihood at the grid search solutions. Our analysis focuses on grid search solutions on the restricted connectivity space C31 = −gC32 + h for simplicity. However, a similar analysis on the bidirectional connections results also in an extreme bias in the connectivity 10 Figure 5: Connections obtained via grid search vs inverse time scale A for σy = 0.0002 (top) and σy = 0.002 (bottom). Left: Mean (solid) and standard deviation (dashed) of grid search solutions on C31 = −gC32 + h over different realizations of the observation noise. All connections are fixed besides the in-connections C32, C31 to node 3. The horizontal black lines are the values of the ground truth connections at C32 = 0.75 and C31 = 0. Right: The averaged minimum of the neg. log-likelihood function for each inverse time scale. The ground truth is given by the model described in section 2.1 but with AGT = 35. The red dot on the right panels represent the minimum of the mean neg. log-likelihood profile. Due to the large variance and the flat neg. log-likelihood the minimum in the bottom case was estimated around A = 40. 11 Figure 6: Standard deviation of grid search solutions for connection C32 depending on the observation noise. The increase in precision needed to obtain reliable connectivity estimates is a direct result of the insensitivity of the BOLD signal to changes in the connectivity. Notice the log-scaling of both axes. estimates due to the wrong time scale. This can be appreciated by inspecting the neg. log-likelihood landscape of a bidirectional connection with a mismatch in A (not shown). As before, if the data is generated by a fast process but one assumes a slow process, the landscape is distorted and barely changes across different noise realizations and values of σy. The distorted landscape features a minimum around (C32, C23) = (1.3,−1.3), although the data was generated using (C32, C23) = (0.75, 0). Finally, we study the variance of the connectivity estimates as a function of the observation noise σy. For this, data is generated with AGT = 1, 5 and 35 and the brute force search is performed without model mismatch. As expected from figure 5, the mean connectivity remains close to the ground truth values for all noise levels (not shown). Figure 6 shows the standard deviation of the grid search connections C32. We observe that the noise level allowed for connectivity estimation depends on the inverse time scale of the neuronal dynamics. There are two effects here. First and most intuitive, the variance of the grid search connections increases with the observation noise. Notice the linear increase for large values in this log-log-graph. Second but more important, fast time scales in the neuronal signal exacerbate this increase at least an order of magnitude due to the insensitivity of the BOLD signal to changes in the connectivity. Notice how the dependency flattens by lowering the noise, first for A = 1, then for A = 5 and finally for A = 35 in the extreme high precision regime σy ∝ 10−5, where the standard deviation of the grid search solutions for A = 1 and A = 5 collapse. This shows that -- even in this idealized case -- to resolve connections when the neuronal time scale is around 30 ms we require at most σy < 10−3. 4 Discussion In this paper we present a method to obtain point estimates of effective connectivity from fMRI time series. This method is based on the Expectation-Maximization algorithm with the E-step being performed by APIS on the same biologically plausible model used in DCM. We show that APIS-EM obtains robust estimations independent of random initializations of the connectivity matrix. In addition, we analyze the effect of the neuronal time scale on estimates of the effective connectivity. This analysis shows that if the underlying neuronal dynamics are fast, the connectivity estimates will be extremely biased if slow processes are assumed in the reconstruction. Hence, jointly estimating the neuronal time scale and the connectivity is important. However, due to the high insensitivity of the BOLD signal for faster neuronal dynamics, this becomes increasingly more demanding and requires higher quality data. We conclude that robust reconstruction of the connectivity between ROIs including the directionality requires much higher precision in the data when assuming a realistic (fast) time scale for neuronal dynamics. In addition, 12 if a slow time scale is assumed (as in [8, 9]) the inferred connectivity can be completely wrong if the generative neuronal process is fast. The adaptive importance sampler APIS is an efficient tool to compute posterior estimates over the hidden processes. Importantly, the sampling step is easily parallelizable such that one can lower the computational burden involved in integrating the system of non-linear differential equations while having a large amount of samples to ensure reliable estimates. Furthermore, APIS and its generalization based on PICE [?], are flexible frameworks that allow an easy adap- tation of the M-step to learn the hemodynamic parameters by including noise in the hemodynamic degrees of freedom. Hence, instead of an m-dimensional controller, APIS would learn a 5m-dimensional controller. Although the computational burden increases in this case, we find empirically that the fully controlled version behaves more stable in terms of the effective sample size. Nevertheless, for simplicity we focused here on the reconstruction of the neural network and assumed the hemodynamic system is known. Hence, to maintain a lower computational cost, we chose the simplified, under-actuated setting. Since the proposed scheme obtains robust estimates from randomly initialized connectivity matrices, it can be considered as a gradient-based exploratory procedure able to change the initial connectivity to find better fitting models. This addresses some of the concerns raised in [17]. Naturally, better initializations improve performance. A natural choice is the symmetric connectivity obtained from methods known to infer the right undirected network, for instance partial correlation or regularized inverse covariance. Moreover, the addition of modulatory inputs or the consideration of more complex neuronal systems, e.g. firing rate models with non-linear activation functions, is straightforward. Furthermore, extensions of the proposed method to obtain Gaussian approximations of the posterior over all parameters similar to DCM can be worked out. This makes the combination of APIS and EM a flexible and efficient alternative to DCMs. Finally, a word of caution. The bias in the connectivity estimates assuming the wrong time scale has significant consequences for the estimation of the hemodynamic parameters because the resulting BOLD signal is shifted roughly a second with respect to the ground truth. If both sets of parameters are jointly estimated, this shift will bias the hemodynamic parameters as well. On the contrary, fast neuronal dynamics make the delays being completely determined by the hemodynamics. This might lower the "interference" between delays caused by the neuronal time scale and the ones caused by the hemodynamics, possibly increasing identifiability of the hemodynamic system. Hence, it is important to clarify the temporal scale of the underlying processes, specially because this assumption has a major impact on the required quality of the data to obtain correct directed networks. Acknowledgements This work was supported by the European Commission through the FP7 Marie Curie Initial Training Network 289146, NETT: Neural Engineering Transformative Technologies. Appendix: Validation of Grid Search on z∗-Line We validate the grid search approach by studying the neg. log-likelihood landscape on (C31, C32) for two extreme cases with and without the correct inverse time scale. We show graphically that when there is a mismatch in A, the bias caused by the restriction on the z∗-line is small compared to the bias caused by the wrong time scale. For the validation, data was generated using the chain network with AGT = 35 and two different noise levels σy = 2 · 10−3 and σy = 2 · 10−4. Figure 7 left panel shows the neg. log-likelihood on (C31, C32) for σy = 0.002 estimated using the correct time scale. Although the high noise level 'washes out' the z∗-line (magenta), there is a clear direction along this line where the values are smallest. Perpendicular to this direction the value increases. Thus, minima lie in this case on the line. Due to the high noise level, the neg. log-likelihood along this line is flat. For lower noise levels, the valley becomes more and more narrow and the profile of the neg. log-likelihood on this line shows a clear minimum. On the right panel of figure 7, we consider more precise data generated with σy = 2 · 10−4 but compute the neg. log-likelihood with a model mismatch in the time scale (A = 1). This distorts the landscape such that the low valued region around the z∗-line transforms into an elliptic shaped valley with the minimum (dot #2) far away from the ground truth (dot #1). Notice how the major axis is parallel to the z∗-line and, although the minimum does not lie exactly on the line, the error is small compare to the bias caused by the wrong time scale. Thus, restricting the search to the z∗-line is a good approximation for the purpose of studying the grid search connections as a function of A. 13 Figure 7: Validating the restriction of the search to the z∗-line (magenta line). Negative log-likelihood in log-scale for two different cases of the parameters A and σy. The upper most data point gives the position of the ground truth. Left: No mismatch in the time scale of the model (A = AGT = 35 Hz) and σy = 0.002. Notice the direction of the valley along the z∗-line. Right: There is a mismatch between the ground truth time scale (AGT = 35 Hz) and the model (A = 1 Hz). Although the noise is small σy = 2· 10−4, there is a wide valley in the form of an ellipse. Notice how the major axis is parallel to the z∗-line. The minimum (lower data point) falls in this case outside the line but the bias introduced by disregarding the perpendicular direction is small compared to the overall bias caused by the wrong time scale. References [1] Michael Breakspear. Dynamic and stochastic models of neuroimaging data: A comment on lohmann et al. Neuroimage, 75:270 -- 274, 2013. [2] Richard B Buxton, Eric C Wong, and Lawrence R Frank. Dynamics of blood flow and oxygenation changes during brain activation: the balloon model. Magnetic resonance in medicine, 39(6):855 -- 864, 1998. [3] Arthur P Dempster, Nan M Laird, and Donald B Rubin. Maximum likelihood from incomplete data via the em algorithm. Journal of the royal statistical society. Series B (methodological), pages 1 -- 38, 1977. [4] Karl Friston, Jean Daunizeau, and Klaas Enno Stephan. Model selection and gobbledygook: Response to lohmann et al. Neuroimage, 75:275 -- 278, 2013. [5] Karl Friston, Rosalyn Moran, and Anil K Seth. Analysing connectivity with granger causality and dynamic causal modelling. Current opinion in neurobiology, 23(2):172 -- 178, 2013. [6] Karl Friston, Klaas Stephan, Baojuan Li, and Jean Daunizeau. Generalised filtering. Mathematical Problems in Engineering, 2010, 2010. [7] Karl J Friston. Functional and effective connectivity in neuroimaging: a synthesis. Human brain mapping, 2(1-2):56 -- 78, 1994. [8] Karl J Friston, Lee Harrison, and Will Penny. Dynamic causal modelling. Neuroimage, 19(4):1273 -- 1302, 2003. [9] Karl J Friston, Baojuan Li, Jean Daunizeau, and Klaas E Stephan. Network discovery with dcm. Neuroimage, 56(3):1202 -- 1221, 2011. [10] Karl J Friston, Andrea Mechelli, Robert Turner, and Cathy J Price. Nonlinear responses in fmri: the balloon model, volterra kernels, and other hemodynamics. NeuroImage, 12(4):466 -- 477, 2000. [11] Igor Vladimirovich Girsanov. On transforming a certain class of stochastic processes by absolutely continuous substitution of measures. Theory of Probability & Its Applications, 5(3):285 -- 301, 1960. [12] Rainer Goebel, Alard Roebroeck, Dae-Shik Kim, and Elia Formisano. Investigating directed cortical interac- tions in time-resolved fmri data using vector autoregressive modeling and granger causality mapping. Magnetic resonance imaging, 21(10):1251 -- 1261, 2003. 14 [13] Clive WJ Granger. Investigating causal relations by econometric models and cross-spectral methods. Econo- metrica: Journal of the Econometric Society, pages 424 -- 438, 1969. [14] Pearl Judea. Causality: models, reasoning, and inference. Cambridge University Press. ISBN 0, 521(77362):8, 2000. [15] Hilbert J Kappen. Linear theory for control of nonlinear stochastic systems. Physical review letters, 95(20):200201, 2005. [16] Hilbert J Kappen, Vicenç Gómez, and Manfred Opper. Optimal control as a graphical model inference problem. Machine learning, 87(2):159 -- 182, 2012. [17] Gabriele Lohmann, Kerstin Erfurth, Karsten Müller, and Robert Turner. Critical comments on dynamic causal modelling. Neuroimage, 59(3):2322 -- 2329, 2012. [18] Joseph B Mandeville, John JA Marota, C Ayata, Greg Zaharchuk, Michael A Moskowitz, Bruce R Rosen, and Robert M Weisskoff. Evidence of a cerebrovascular postarteriole windkessel with delayed compliance. Journal of Cerebral Blood Flow & Metabolism, 19(6):679 -- 689, 1999. [19] Radford M Neal and Geoffrey E Hinton. A view of the em algorithm that justifies incremental, sparse, and other variants. In Learning in graphical models, pages 355 -- 368. Springer, 1998. [20] Rajan S Patel, F DuBois Bowman, and James K Rilling. A bayesian approach to determining connectivity of the human brain. Human brain mapping, 27(3):267 -- 276, 2006. [21] Will Penny, Zoubin Ghahramani, and Karl Friston. Bilinear dynamical systems. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 360(1457):983 -- 993, 2005. [22] Joseph D Ramsey, Stephen José Hanson, Catherine Hanson, Yaroslav O Halchenko, Russell A Poldrack, and Clark Glymour. Six problems for causal inference from fmri. neuroimage, 49(2):1545 -- 1558, 2010. [23] Hans-Christian Ruiz and Hilbert J Kappen. Particle smoothing for hidden diffusion processes: Adaptive path integral smoother. IEEE Transactions on Signal Processing, 65(12):3191 -- 3203, 2017. [24] H.-Ch. Ruiz Euler and H.J. Kappen. Nonlinear deconvolution by sampling a biophysically plausible hemody- namic model. 2017. [25] Srikanth Ryali, Kaustubh Supekar, Tianwen Chen, and Vinod Menon. Multivariate dynamical systems models for estimating causal interactions in fmri. Neuroimage, 54(2):807 -- 823, 2011. [26] Simo Särkkä. Bayesian filtering and smoothing, volume 3. Cambridge University Press, 2013. [27] Thomas B Schön, Adrian Wills, and Brett Ninness. System identification of nonlinear state-space models. Automatica, 47(1):39 -- 49, 2011. [28] Stephen M Smith, Karla L Miller, Gholamreza Salimi-Khorshidi, Matthew Webster, Christian F Beckmann, Thomas E Nichols, Joseph D Ramsey, and Mark W Woolrich. Network modelling methods for fmri. Neuroimage, 54(2):875 -- 891, 2011. [29] Klaas Enno Stephan, Lars Kasper, Lee M Harrison, Jean Daunizeau, Hanneke EM den Ouden, Michael Break- spear, and Karl J Friston. Nonlinear dynamic causal models for fmri. Neuroimage, 42(2):649 -- 662, 2008. 15
1606.08232
1
1606
2016-06-27T12:15:39
Information integration from distributed threshold-based interactions
[ "q-bio.NC", "cond-mat.stat-mech" ]
We consider a collection of distributed units that interact with one another through the sending of messages. Each message carries a positive ($+1$) or negative ($-1$) tag and causes the receiving unit to send out messages as a function of the tags it has received and a threshold. This simple model abstracts some of the essential characteristics of several systems used in the field of artificial intelligence, and also of biological systems epitomized by the brain. We study the integration of information inside a temporal window as the model's dynamics unfolds. We quantify information integration by the total correlation, relative to the window's duration ($w$), of a set of random variables valued as a function of message arrival. Total correlation refers to the rise of information gain above and beyond that which the units already achieve individually, being therefore related to consciousness studies in some models. We report on extensive computational experiments that explore the interrelations of the model's parameters (two probabilities and the threshold), highlighting relevant scenarios of message traffic and how they impact the behavior of total correlation as a function of $w$. We find that total correlation can occur at significant fractions of the maximum possible value and provide semi-analytical results on the message-traffic characteristics associated with values of $w$ for which it peaks. We then reinterpret the model's parameters in terms of the current best estimates of some quantities pertaining to cortical structure and dynamics. We find the resulting possibilities for best values of $w$ to be well aligned with the time frames within which percepts are thought to be processed and eventually rendered conscious.
q-bio.NC
q-bio
Information Integration from Distributed Threshold-Based Interactions Valmir C. Barbosa∗ Programa de Engenharia de Sistemas e Computa¸cao, COPPE Universidade Federal do Rio de Janeiro Caixa Postal 68511 21941-972 Rio de Janeiro - RJ, Brazil Abstract We consider a collection of distributed units that interact with one another through the sending of messages. Each message carries a pos- itive (+1) or negative (−1) tag and causes the receiving unit to send out messages as a function of the tags it has received and a threshold. This simple model abstracts some of the essential characteristics of sev- eral systems used in the field of artificial intelligence, and also of biological systems epitomized by the brain. We study the integration of information inside a temporal window as the model's dynamics unfolds. We quantify information integration by the total correlation, relative to the window's duration (w), of a set of random variables valued as a function of message arrival. Total correlation refers to the rise of information gain above and beyond that which the units already achieve individually, being therefore related to consciousness studies in some models. We report on extensive computational experiments that explore the interrelations of the model's parameters (two probabilities and the threshold), highlighting relevant scenarios of message traffic and how they impact the behavior of total correlation as a function of w. We find that total correlation can occur at significant fractions of the maximum possible value and provide semi- analytical results on the message-traffic characteristics associated with values of w for which it peaks. We then reinterpret the model's parame- ters in terms of the current best estimates of some quantities pertaining to cortical structure and dynamics. We find the resulting possibilities for best values of w to be well aligned with the time frames within which percepts are thought to be processed and eventually rendered conscious. Keywords: Threshold-based system, Information integration, Total cor- relation, Cortical dynamics, Percept, Consciousness. ∗[email protected]. 1 1 Introduction A threshold-based system is a collection of loosely coupled units, each charac- terized by a state function that depends on how the various inputs to the unit relate to a threshold parameter. The coupling in question refers to how the units interrelate, which is by each unit communicating its state to some of the other units whenever that state changes. Given a set of timing assumptions and how they relate to the exchange of states among units as well as to state updates, each individual unit processes inputs (states communicated to it by other units) and produces a threshold-dependent output (its own new state, which gets communicated to other units). The quintessential threshold-based system is undoubtedly the brain, where each neuron's state function determines whether an action potential is to be fired down the neuron's axon. This depends on how the combined action potentials the neuron perceives through the synapses connecting other neurons' axons to its dendrites (its synaptic potentials) relate to its threshold potential [1]. The greatly simplified model of the natural neuron known as the McCulloch- Pitts neuron [2], introduced over seventy years ago, retained this fundamental property of being threshold-based, and so did generalizations thereof such as generalized Petri nets [3] and threshold automata [4]. In fact, this holds for much of the descent from the McCulloch-Pitts neuron, which has extended through the present day in a succession of ever more influential dynamical systems. Such descent includes the essentially deterministic Hopfield networks of the 1980s [5,6] and moves on through generalizations of those networks' Ising-model type of energy function and the associated need for stochastic sampling. The resulting networks include the so-called Boltzmann machines [7] and Bayesian networks [8, 9], as well as the more general Markov (or Gibbs) Random Fields [10 -- 12] and several of the probabilistic graphical models based on them [13]. A measure of the eventual success of such networks can be gained by considering, for example, the restricted form of Boltzmann machines [14, 15] used in the construction of deep belief networks [16], as well as some of the other deep networks that have led to landmark successes in the field of artificial intelligence recently [17 -- 19]. Our interest in this paper is the study of how information gets integrated as the dynamics of a threshold-based system is played out. The meaning we attach to the term information integration is similar to the one we used previ- ously in other contexts [20, 21]. Given a certain amount of time w and a set of random variables, each related to the firing activity of each of the system's units inside a temporal window of duration w, we quantify integrated informa- tion as the amount of information the system generates as a whole (relative to a global state of maximum entropy) beyond that which accounts for the aggre- gated information the units generate individually (now relative to local states of maximum entropy). This surplus of information is known as total correla- tion [22] and is fundamentally dependent on how the units interact with one another. 2 Our understanding of information integration, therefore, lies in between those used by approaches that seek it in the synchronization of input/output signals (see [23], for example) and those that share our view but would con- sider not just the whole and the individual units but all partitions in between as well [24]. By virtue of this, we remain aligned with the latter theory by ac- knowledging that information gets integrated only when it is generated by the whole in excess of the total its parts can generate individually. On the other hand, by sticking with total correlation as an information-theoretic quantity that requires only two partitions of the set of units to be considered (one that is fully cohesive and another that is maximally disjointed), we ensure tractability way beyond that of the all-partitions theory. We conduct all our study on a simple model of threshold-based systems. In this model, the units are placed inside a cube and exchange messages whose delivery depends on their propagation speed and the Euclidean distance between sender and receiver. Every message is tagged, and upon reaching its destination, its tag is used to move a local accumulator either toward or away from the threshold. Reaching the threshold makes the unit send out messages and the accumulator is reset. There are three parameters in the model. Two of them are probabilities (that a message is tagged so that the accumulator at the destination gets decreased upon its arrival, and that a unit sends a message to each of the other units), the other being the value of the threshold. Parameter values are the same for all units. This model is by no means offered as an accurate representation of any par- ticular threshold-based system, but nevertheless summarizes some key aspects of such systems through its relatively few parameters. In particular, it gives rise to three possible expected global regimes of message traffic. One of them is perfectly balanced, in the sense that on average as much traffic reaches the units as leaves them. In this case, message traffic is sustained indefinitely. In each of the other two regimes, by contrast, either more traffic reaches the units than leaves them, or conversely. Message traffic dies out in the former of these two (unless the units receive further external input) but grows indefinitely in the latter one. We find that information integration is guaranteed to occur at high levels for some window durations whenever message traffic is sustained at the perfect- balance level or grows. We also find that this happens nearly independently of parameter variations. On the other hand, we also find that information inte- gration is strongly dependent on the model's parameters, with significant levels occurring only for some combinations, whenever message traffic is imbalanced toward the side that prevents it from being sustained. Here we once again turn to the brain, whose cortical activity is in many accounts characterized as tending to be sparse [25, 26], as an emblematic example. We proceed as follows. Our message-passing model is laid out in Section 2, where its geometry and distributed algorithm are detailed and the question of message imbalance is introduced. An account of our use of total correlation is given in Section 3, followed by our methodology in Section 4. Results, discussion, and conclusion follow, respectively in Sections 5, 6, and 7. 3 2 Model Our system model comprises a structural component and an algorithmic one. The two are described in what follows, along with some analysis of how they interrelate. 2.1 Underlying geometry For 1 ≤ D ≤ 3, our model is based on N simple processing units, henceforth referred to as nodes, each one placed at a fixed position inside the D-dimensional cube of side ℓ. The position of node i has coordinates X (1) , so the Euclidean distance between nodes i and j is , . . . , X (D) i i δij = q(X (1) i − X (1) j )2 + · · · + (X (D) i − X (D) j )2. (1) We assume that nodes can communicate with one another by sending messages that propagate at the fixed speed σ on a straight line. The delay incurred by a message sent between nodes i and j in either direction is therefore δij /σ. Our computational experiments will all be such that the N nodes are placed in the D-dimensional cube uniformly at random. In this case, and in the limit of infinite N , the expected distance between two randomly chosen nodes i and j is given by ℓ = (cid:18) 1 ∆(D) ℓ(cid:19)2DZ ℓ 0 · · ·Z ℓ 0 δijdX (1) i · · · dX (D) i dX (1) j · · · dX (D) j , (2) where 1/ℓ is the probability density for each of the 2D variables. Letting X (d) ℓx(d) k in this equation for k ∈ {i, j} and d ∈ {1, . . . , D} yields k = ∆(D) ℓ = ℓ2D+1 ℓ2D Z 1 0 · · ·Z 1 0 δijdx(1) i · · · dx(D) i dx(1) j · · · dx(D) j , where δij now has x's in place of the X's. We then have ∆(D) ℓ = ℓ∆(D) 1 , (3) (4) with the expected distances in the unit cube for the numbers of dimensions of interest being well known: ∆(1) 1 ≈ 0.6617 [28]. 1 ≈ 0.5214 [27], and ∆(3) 1 = 1/3, ∆(2) In addition to expected distances, the associated variances will also at one point be useful. Analytical expressions for most of them seem to have remained unknown thus far, but the underlying probability densities have been found to be more concentrated around the expected values given above as D grows [29]. That is, variance is greatest for D = 1. 4 2.2 Network algorithmics We view the N nodes as running an asynchronous message-passing algorithm collectively. By asynchronous we mean that each node remains idle until a message arrives. When this happens, the arriving message is processed by the node, which may result in messages being sent out as well. Such a purely reactive stance on the part of the nodes requires at least one node to send out at least one message without having received one, for startup. We assume that this is done by all nodes initially, after which they start behaving reactively. We assume that each message carries a signed-unit tag (i.e., either +1 or −1) with it. The specific tag to go with the message is chosen probabilistically by its sender at send time, with −1 being chosen with probability p−. The processing done by node i upon arrival of a message is the heart of the system's thresholding nature and involves manipulating an accumulator Ai, initially equal to 0, to which every tag received is added (unless Ai = 0 and the tag is −1, in which case Ai remains unchanged). Whenever Ai reaches a preestablished integer value τ > 0, node i sends out messages of its own and Ai is reset to 0. Thus, the integer τ acts as a threshold governing the sending of messages by (the firing of) node i. The values of p− and τ are the same for all nodes. It follows from this simple rule that the value of Ai is perpetually confined to the interval [0, τ ]. The expected number of messages that node i has to receive in order for Ai to be increased all the way from 0 to τ is the node's expected number of message arrivals between firings, henceforth denoted by µ = µ(p−, τ ). The value of µ can be calculated easily once we recognize that µ is simply the expected number of steps for the following discrete-time Markov chain to reach state τ having started at state 0. The chain has states 0, . . . , τ and transition probability prs, from state r to state s, given by 1 − p− if s = r + 1; p− 1 0 if s + 1 = r < τ or r = s = 0; if r = s = τ ; otherwise. prs =   It is easily solved and yields µ = p− (1 − 2p−)2 (cid:20)(cid:18) p− 1 − p−(cid:19)τ − 1(cid:21) + τ 1 − 2p− (5) (6) for p− 6= 0.5 [30, page 348].1 The sending of messages when a node fires is based on another parameter, p→, which is the probability with which the node sends a message to each of the other N − 1 nodes. It follows that the expected number of messages that get sent out is (N − 1)p→. The value of p→ is the same for all nodes as well. 1For p− = 0.5 we have µ = τ (τ + 1) [30, page 349], but this holds for none of our computational experiments. 5 2.3 Local imbalance and global message traffic At node i, a balance exists between message output and message input when the expected number of messages sent out at each firing is the same as the expected number of messages received between two successive firings. That is, message traffic is locally balanced when (N − 1)p→ = µ. It is locally imbalanced otherwise, which can be quantified by the difference I, defined to be I = (N − 1)p→ µ − 1. (7) Given I, clearly the instantaneous density of global message output at time t, denoted by M (t), is expected to remain constant with t if I = 0, or to decrease or increase exponentially with t depending on whether I < 0 or I > 0, respectively. This behavior is described by dM (t) dt = I T0 M (t), (8) where the T0 participating in the time constant T0/I is some fundamental amount of time related to the system's geometry and kinetics. In Section 5, we provide empirical evidence that T0 is the expected delay undergone by a message, given by T0 = ∆(D) ℓ σ . Equation (8) is of immediate solution, yielding where M (t) = M0eIt/T0, M0 = M (0) = N (N − 1)p→ (9) (10) (11) is the expected number of messages that all nodes, collectively, send out initially. Similarly, the cumulative global message output inside a temporal window of duration w starting at time t is Mt(w) = Z t+w t M (t)dt = M0T0 I (cid:16)eI(t+w)/T0 − eIt/T0(cid:17) . (12) In the case of locally balanced message traffic (I = 0), this expression is easily seen to yield Mt(w) = M0w, (13) therefore independent of t. Otherwise, Mt(w) either decreases or increases ex- ponentially with t, depending respectively on whether I < 0 or I > 0. 6 2.4 Graph-theoretic interpretations The model described so far in Section 2 can be regarded as a directed geometric graph, that is, a graph whose nodes are positioned in some region of interest (the D-dimensional cube of side ℓ) and whose edges are directed. It is moreover a complete graph without self-loops, in the sense that an edge exists directed from every node i to every node j 6= i. Our use of the model in the sequel will require the nodes to be positioned at random before each new run of the distributed algorithm of Section 2.2, so an alternative interpretation that one might wish to consider views the model as a variation of the traditional random geometric graph [31]. In this variation, an edge exists directed from i to j 6= i with fixed probability p→, independently of any other node pair. That is, aside from node positioning the graph underlying our model is an Erdos-R´enyi random graph [32] as extended to the directed case [33]. This interpretation is somewhat loose, though, because it requires that we view each individual run of the algorithm as being equivalent to several runs on independent instances of the underlying random graph, with multiple further runs serving to validate any statistics that one may come up with at the end. This is hard to justify, however, particularly when one considers the nonlinear- ities characterizing the quantities we will average over all runs of the algorithm (see Section 3). Even so, interpreting our model in terms of random graphs remains tantalizing in some contexts. For example, it allows the parameter p− to be regarded as the fraction of a node's in-neighbors from which messages with negative tags are received. In the context of networks of the brain at the neuronal level, for example, an abstract rendering of the fraction of neurons that are inhibitory is obtained (see Section 6.1). 3 Total correlation We use the total correlation of N random variables [22], each corresponding to one of the N nodes, as a measure of information integration. Each of these variables is relative to a temporal window of fixed duration w, the variable corresponding to node i being denoted by X (w) and taking up values from the set {0, 1}. The intended semantics is that X (w) = 1 if and only if node i receives at least one message in a time interval of duration w. We also use the shorthands X(w) and x to denote the sequence of variables (X (w) , . . . , X (w) N ) and the sequence of values (x1, . . . , xN ), respectively. Thus, X(w) = x stands for the joint valuation X (w) 1 = x1, . . . , X (w) N = xN . 1 i i Given the marginal Shannon entropy of each variable X (w) , i H(X (w) i ) = − Xx∈{0,1} Pr(X (w) i = x) log2 Pr(X (w) i = x), (14) 7 and the joint Shannon entropy H(X(w)) = − Xx∈{0,1}N Pr(X(w) = x) log2 Pr(X(w) = x), (15) the total correlation of the N variables given w is defined as2 C(X(w)) = N Xi=1 H(X (w) i ) − H(X(w)). (16) To see the significance of this definition in our context, consider the flat joint probability mass function, Pr(X(w) = x) = 2−N for all x ∈ {0, 1}N. This mass function entails maximum uncertainty of the variables' values, hence the maximum possible value of the joint entropy, H(X(w)) = N . It also implies flat marginals, Pr(X (w) i = x) = 0.5 for all x ∈ {0, 1} and all i, and again the maximum possible value of each marginal entropy, H(X (w) ) = 1. The difference from the actual joint entropy H(X(w)) to its maximum reflects a reduction of uncertainty, or an information gain, i G(X(w)) = H(X(w)) − H(X(w)) = N − H(X(w)), the same holding for each of the marginals, G(X (w) i ) = H(X (w) i ) − H(X (w) i ) = 1 − H(X (w) i ). (17) (18) Thus, it is possible to rewrite the expression for C(X(w)) in such a way that G(X(w)) = C(X(w)) + N Xi=1 G(X (w) i ). (19) That is, the total correlation of all N variables is the information gain that surpasses their combined individual gains. This surplus is zero if and only if the variables are independent of one another, that is, precisely when Pr(X(w) = x) = i = xi) for all x ∈ {0, 1}N , since in this case we have H(X(w)) = ). It is strictly positive otherwise, with a maximum possible value i i=1 Pr(X (w) QN i=1 H(X (w) PN of N − 1. Achieving this maximum requires a joint probability mass function assigning no mass to any but two members of {0, 1}N , say x and y, and moreover that these two be equiprobable (i.e., Pr(X(w) = x) = Pr(X(w) = y) = 0.5) and complementary to each other (i.e., xi + yi = 1 for every i). Referring back to the intended meaning of the N random variables, total correlation is maximized in those runs of the distributed algorithm of Section 2.2 for which a partition (X1, X2) of the set {X (w) N } exists with the following two properties. , . . . , X (w) 1 2The N = 2 case of this formula coincides with that for mutual information, but one is to note that in the general case the two formulas are completely different [34]. 8 First, no matter which particular window of duration w we concentrate on, the set of nodes that receive at least one message inside the window is either X1 or X2. Second, the first property holds with X1 for exactly half such windows. While these are exceedingly stringent conditions both spatially and tempo- rally, perhaps implying that values of total correlation equal to or near N −1 are practically unachievable, they serve to delineate those scenarios with a chance of generating substantial amounts of total correlation. Specifically, such scenarios will on average have a pattern of global message traffic, inside a window, that is neither too sparse nor too dense. Furthermore, sustaining such an amount of to- tal correlation as time elapses will also require traffic patterns that deviate only negligibly from the ones yielding the average, possibly entailing some variability on window sizes. Our methodology to track and validate values of w leading to noteworthy total correlation is described next. It involves computational ex- periments for a variety of values for w, and also gauging the cumulative global message output Mt(w) that results from each experiment against a function of the reference number of messages and reference delay embodied in M0 and T0, respectively. 4 Methods Our results are based on running the distributed algorithm of Section 2.2 for a fixed geometry (i.e., fixed number of dimensions D ∈ {1, 2, 3}, fixed value of the cube side ℓ, and fixed positioning of N nodes in the cube) and a fixed set of values for the parameters (p−, τ , and p→). Each run of the algorithm terminates either when no more message is in transit (so none will ever be thenceforth, given the algorithm's reactive nature) or when a preestablished maximum number of messages in transit has been reached, whichever comes first. Imposing the latter upper bound is important because it serves to size the data structures where messages in transit are kept for later processing. Node positioning is achieved uniformly at random, so multiple runs are needed for each fixed configuration (D, ℓ, N, p−, τ, p→). Each run leaves a trace of all events taking place as it unfolds, each event referring to the arrival of a message at a node and comprising the node's identification and the message's arrival time. A series of values for w is then considered and for each w each of the traces is analyzed, yielding the total correlation produced by the corresponding run. The average total correlation over all the runs is then reported. Following our discussion at the end of Section 3, for each value of w we gauge Eq. (12) against the approximation to it given by M0(T0) = M0T0, according to which M0 messages get sent at time t = 0 and received at time T0. We do this by postulating a proportionality constant α > 0 between them, that is, by assuming Mt(w) = αM0T0. (20) Doing this allows us to express w as a function of α for each t and, whenever possible, to characterize traffic regimes giving rise to substantial amounts of total correlation. 9 4.1 Supporting analysis We denote the value of t upon termination of a run by T and the total number of messages sent by M . An approximation to Eq. (12) similar to the one above can be used to relate T and M as M0(T ) = M T0, whose right-hand side quantifies what would be expected to happen if all M messages were sent at time t = 0 and received at time T0. This leads to T = T0 I ln(cid:18)1 + M I M0 (cid:19) . Solving Eq. (20) for w = w(t) given α yields w(t) = T0 I ln(cid:16)1 + αIe−It/T0(cid:17) , whose value for t = 0 is the duration of the first window, w0 = w(0) = T0 I ln(1 + αI). (21) (22) (23) As for the duration of the last window, which we denote by w∗, it can likewise be found by solving Eq. (20) for w, now letting the window's start time be t = T − w, then letting w∗ = w. We obtain w∗ = T0 I ln(cid:18) 1 + M I/M0 1 + (M/M0 − α)I(cid:19) . The average window duration between time t = 0 and time T − w∗ = T0 I ln(cid:18)1 +(cid:18) M M0 − α(cid:19) I(cid:19) , denoted by ¯w, is also of interest and comes from the indefinite integral I (cid:19)2 f (t) = Z w(t)dt = (cid:18) T0 Li2(cid:16)−αIe−It/T0(cid:17) , (24) (25) (26) where Li2(z) = Pk≥1 zk/k2 is the dilogarithm of z. This given, we obtain ¯w = f (T − w∗) − f (0) T − w∗ = T0 I Li2(−αI/[1 + (M/M0 − α) I]) − Li2(−αI) ln(1 + (M/M0 − α) I) . (27) While for I = 0 we have T = M T0/M0 and w0 = ¯w = w∗ = αT0, (28) for I 6= 0 everything depends on the sign of I. If I < 0, then we need M/M0 < −1/I in order for T to be well defined. Moreover, we have 0 < w0 < ¯w < w∗ < T, (29) 10 Table 1: Parameter values for each of settings I -- IV and each of the three possibil- ities for the value of I in relation to 0. Values of p→ are given by p→ = µ/(N −1) for I = 0, as per Eq. (7). Setting N τ p− I II III IV 500 300 500 500 0.30 0.30 0.18 0.30 5 5 5 10 I < 0 0.01 0.01 0.01 0.01 p→ I = 0 0.02135 0.03563 0.01478 0.04634 I > 0 0.06 0.06 0.06 0.06 where the first inequality holds if α < M/M0, this being necessary and sufficient for the last inequality to hold as well. For I > 0, on the other hand, T is always well defined and we get T > w0 > ¯w > w∗ > 0. (30) In this case, the constraint α < M/M0 is necessary and sufficient for both the first and the last inequality to hold. 4.2 Computational experiments We organize our computational experiments into settings, numbered I -- IV, each comprising all configurations (D, ℓ, N, p−, τ, p→) for which N , p−, and τ are fixed. In each setting, there are three possibilities for the value of p→, one ensuring I < 0 (p→ = 0.01), one for I = 0 (extracted from Eq. (7)), and one for I > 0 (p→ = 0.06). The four settings are summarized in Table 1. Each of settings II -- IV is derived from setting I by a change in the value of N , p−, or τ , respectively. Each setting entails 2 100 runs of the algorithm for each p→ value, this total comprising 100 independent trials for each combination of D ∈ {1, 2, 3} and ℓ ∈ {10−3, 10−2, . . . , 103}. Each run starts by placing the N nodes anew, uniformly at random, and proceeds therefrom subject to the further indeterminacies that characterize the algorithm through the probabilities p− and p→. Messages are assumed to travel at the speed σ = 1. Each of the 100 traces resulting from the same configuration is then analyzed and the corresponding value of C(X(w)) is computed, the average over the traces being reported at the end. This is done for each w ∈ {2−33, 2−32.5, . . . , 210}. One last important aspect of a run has to do with the determination of the maximum number of messages ever allowed to be in transit as the run unfolds. We determine this number via the formula 1 000N µ, where 1 000µ can be in- terpreted as the expected number of messages a node must receive if it is to fire 1 000 times during the run. This latter number is in many ways arbitrary, though, having to do only with ensuring that the computational resources re- quired by all runs remain manageable. As a result, taken as a whole the 25 200 runs have generated a total of about 394 gigabytes of trace in compressed format. Moreover, processing these data for the determination of the various C(X(w)) 11 averages has required several weeks of computation on an Intel Xeon E5-1650 core clocked at 3.2 GHz and having exclusive access to 30 gigabytes of RAM. 5 Results The average value of the total correlation C(X(w)) resulting from the compu- tational experiments described in Section 4.2 are presented in normalized form (i.e., relative to the maximum N − 1) in Figures 1, 2, and 3, respectively for the I < 0 cases, the I = 0 cases, and the I > 0 cases. Each figure comprises four panels, numbered I -- IV to match the four settings of Table 1. Each panel contains 21 plots, each corresponding to one of the possible variations in the number of dimensions D and in the cube side ℓ. Each plot in the three figures is given against the rescaled version of the window size w given by w/T0, where we recall from Eq. (9) that T0 is the delay a message incurs when traversing the distance ∆(D) at speed σ. That is, the abscissas in Figures 1 -- 3 are all relative to the geometric and kinetic underpinnings to which each plot refers. This rescaling reveals that, for each fixed combination of a setting with an imbalance profile (i.e., for each panel in the three figures), the behavior of the average C(X(w)) is essentially invariant with respect to how the various D and ℓ values are paired. Our choice in Section 2.3 of T0/I as the system's time constant is then clearly justified, as the value of I is fixed in each of the figures' panels. Moreover, such invariance also backs up our choice of σ = 1 for all simulations (see Section 4.2), since the choice of any other value would simply alter the rescaling factors. ℓ Nevertheless, there are signs in Figures 3(I) and (III) that such invariance may not hold quite as well when D = 1. We attribute this to the fact that ∆(D) is only an expected distance and as such affects the role of T0 as a rescaling factor differently for each number of dimensions D. In particular, recalling from Section 2.1 that the corresponding variances get larger as D is decreased, it seems clear that what is affecting invariance in the two figures in question is precisely the poor representativeness of ∆(1) . Even so, we see in Figure 3 that this affects setting III much more severely than it does setting I. The reason for this has to do with the value of imbalance I in each case, as we discuss in Section 6. ℓ ℓ In all but the case of Figure 1(IV), the average total correlation starts off at some negligibly small value, then slowly climbs toward a peak along an increase in w/T0 by some orders of magnitude, and finally recedes as w/T0 is made to vary by further orders of magnitude. The plots in each of the panels of Figures 1 -- 3 are given against a backdrop of two vertical lines, the leftmost one marking the smallest w/T0 at which total correlation peaks for some (D, ℓ) pair, the rightmost marking the largest such value. (The smallest w/T0 value in the setting-III panel of Figure 1 does not take into account the cases D = 2, 3 with ℓ = 103, whose peaks seem to occur past w = 210, the largest window used in the simulations, and are therefore unknown.) 12 I n o i t l a e r r o c l a t o t . m r o N II l n o i t a e r r o c l a t o t . m r o N III n o i t l a e r r o c l a t o t . m r o N IV n o i t l a e r r o c l a t o t . m r o N 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 10-4 0 10-5 10-2 10-3 100 Rescaled window size 10-1 101 102 Figure 1: Normalized total correlation (C(X(w))/(N − 1)) against rescaled win- dow size (w/T0) for the I < 0 cases. In the keys to the plots, u = log10 ℓ. Vertical lines mark the smallest and largest w/T0 values given in Table 2 (the smallest such value for setting IV is off range). I -- IV identify settings. 13 I n o i t l a e r r o c l a t o t . m r o N II l n o i t a e r r o c l a t o t . m r o N III n o i t l a e r r o c l a t o t . m r o N IV n o i t l a e r r o c l a t o t . m r o N 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 10-4 0 10-5 10-2 10-3 100 Rescaled window size 10-1 101 102 Figure 2: Normalized total correlation (C(X(w))/(N − 1)) against rescaled win- dow size (w/T0) for the I = 0 cases. In the keys to the plots, u = log10 ℓ. Vertical lines mark the smallest and largest w/T0 values given in Table 2. I -- IV identify settings. 14 I n o i t l a e r r o c l a t o t . m r o N II l n o i t a e r r o c l a t o t . m r o N III n o i t l a e r r o c l a t o t . m r o N IV n o i t l a e r r o c l a t o t . m r o N 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 10-5 10-4 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 1, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 2, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 D = 3, u = -3 u = -2 u = -1 u = 0 u = 1 u = 2 u = 3 101 10-2 10-3 100 Rescaled window size 10-1 102 Figure 3: Normalized total correlation (C(X(w))/(N − 1)) against rescaled win- dow size (w/T0) for the I > 0 cases. In the keys to the plots, u = log10 ℓ. Vertical lines mark the smallest and largest w/T0 values given in Table 2. I -- IV identify settings. 15 The backdrop lines' locations are detailed in Table 2, where the (D, ℓ) giving rise to each one is also shown. The table also contains for each such location a value for α, the proportionality constant that through Eq. (20) is used to relate the global message output Mt(w) inside a size-w window beginning at time t to the reference output M0T0. For I = 0, we have found in Section 4.1 that the w = w(α) necessary for Eq. (20) to hold is time-invariant and given as in Eq. (28), whence α = w/T0. This is reflected in Table 2. For I 6= 0, on the other hand, satisfying Eq. (20) for some fixed α requires w to vary with time according to Eq. (22). This is what one would expect, since the dwindling message output that results from an I < 0 scenario re- quires an ever larger window size to accommodate the amount of traffic given by αM0T0, and likewise the output expansion caused by I > 0 requires pro- gressively smaller window sizes. This dependency on time is reflected in the inequalities of Eqs. (29) and (30), respectively for I < 0 and I > 0, where the earliest window size (w0), as well as the average one ( ¯w) and the latest (w∗) are put in perspective. The values of α given in Table 2 for the I 6= 0 cases come from Eq. (27) by letting each value of w/T0 reported in the table be such that w = ¯w(α). This use of Eq. (27) requires the expected value of M (the number of mes- sages sent in a run) to be estimated from the 100 independent trials for each configuration of interest. We do this by first averaging the number of messages received during each run (let E denote this average) and then recalling that each message received entails an expected number of messages sent equal to (N − 1)p→/µ. Our estimate of the expected M is then3 M = M0 + E(N − 1)p→ µ . (31) An illustration highlighting the use of Eq. (27) on the setting-III cases of Table 2 for I 6= 0 is given in Figure 4, where the plots relative to Eq. (24) rely on the same estimate of the expected M as above. 6 Discussion Even though our notation for total correlation, C(X(w)), may suggest that it is a function solely of the N random variables in X(w), it is in fact a function of the joint probability mass function associated with those variables as well. Therefore, changing the allotment of probability mass to the various points in {0, 1}N must have an impact on the value of C(X(w)) as a matter of principle. Thus, while we know that such variation must never lead total correlation to fall below zero or above N − 1, figuring out the resulting expected value has the 3Our simulator is event-based, each event being the reception of a message. The total we have available at the end of each simulation is that of messages received, not sent, and the two may differ because not every message sent gets received, by virtue of the preestablished maximum number of messages in transit having been reached (see Section 4). We then need this workaround to find the expected M through E. 16 Table 2: Smallest and largest w/T0 at which C(X(w)) peaks for each combination of a setting with a possibility for I relative to 0, all configurations in each combination considered. Each of the two extremes occurs for the configuration to which the D and ℓ values shown refer. For each of the two extremes, the value of α reported alongside it is such that w = ¯w(α). The smallest w/T0 reported for the I < 0, setting-III case does not take into account the peaks for D = 2, 3 with ℓ = 103, which have remained undiscovered by the simulations. 1 7 Imbalance Setting I < 0 I = 0 I > 0 I II III IV I II III IV I II III IV ℓ 1 α 10−3 10−2 102 103 10−3 w/T0 1.47582 0.41743 2.45490 2 × 10−13 Smallest w/T0 for C(X(w)) peak D 3 3 2 2 3 1 2 1 1 1 1 1 0.710 0.116 1.395 0.050 0.522 0.530 0.530 0.586 0.009 0.008 0.009 0.007 0.00129 0.00114 0.00129 0.00101 10−2 10−2 10−1 10−2 10−1 10−3 = α = α = α = α α w/T0 2.17223 0.84853 3.41955 0.18729 Largest w/T0 for C(X(w)) peak D 1 1 3 2 2 3 2 1 3 2 3 3 1.073 0.302 1.917 0.050 0.848 0.774 0.848 0.960 0.015 0.012 0.017 0.010 ℓ 103 10 10 10−2 10−1 103 10−1 102 103 102 10 1 0.00214 0.00170 0.00236 0.00148 = α = α = α = α A i i e z s w o d n w d e a c s e R l B i i e z s w o d n w d e a c s e R l 4.5 4 3.5 3 2.5 2 1.5 1 0.5 0 10-1 10-2 10-3 10-4 10-5 10-6 10-7 0 0.5 1 0 0.004 0.008 Eq. (23) Eq. (27) Eq. (24) 1.5 2 2.5 Eq. (23) Eq. (27) Eq. (24) 0.012 0.016 0.02 α α Figure 4: Rescaled window size (w/T0) against α, as per Eqs. (23), (27), and (24), respectively for w = w0, w = ¯w, and w = w∗. All plots refer to setting III, with either I < 0 (A) or I > 0 (B). The corresponding (α, w/T0) pairs in Table 2 are highlighted as bullets. Plot ranges are given by the inter- vals [0.05, M/M0) in A (to within a difference of 10−4 on the upper side) and [0.001, 0.019] in B. 18 potential of aiding in the interpretation of any particular figure one encounters in some situation of interest. As far as we know, such expected value has so far remained unknown, but resorting to results on the expected value of the joint Shannon entropy (H(X(w)), in our case) [35], it is possible to ascertain that an upper bound on the expected value of C(X(w)) is approximately 0.6 [20]. Once normalized to N − 1, this upper bound would be unnoticeable in any of the panels in Figures 1 -- 3. In fact, it would fall two orders of magnitude below the nearly flat values of Figure 1(IV), which are of the order of 10−1. Clearly, then, in all of settings I -- IV and for all three imbalance scenarios, our model is seen to be promoting the rise of probability mass functions leading to total correlations significantly above the expected value. In most cases, this occurs over ranges for the value of w, the temporal-window duration that provides meaning to the random variables in X(w), spanning several orders of magnitude. Not only this, but in most cases we have found normalized total correlation to peak at substantial levels: between 0.22 and 0.67 in settings I and III when I < 0 (Figure 1); 0.77 and 0.86 in all four settings when I = 0 (Figure 2); and 0.93 and 0.97 in all four settings when I > 0 (Figure 3). The peaks relating to the I = 0 cases (and also those of the I > 0 cases, but to a more limited extent) seem to occur largely independently of the setting in question, that is, regardless of the number of nodes N or the parameters that control firing (the probability p− and the threshold τ ). Understanding this independence comes from focusing on the I = 0 cases, since for I = 0 the value of probability p→ varies from setting to setting as a function of N and µ (therefore a function of p− and τ as well, by Eq. (6)); see Table 1. Setting p→ in this way aims precisely to cause I = 0 and, as a consequence, tends to compensate for any dependency of total-correlation peaks on N , p−, or τ . It also suggests that the value of I, along with that of w/T0, is a major player when it comes to determining how much total correlation can be achieved. However, increasing the value of p→ from those used to ensure I = 0 to p→ = 0.06, thus ensuring I > 0, preserves nearly the same independence of peak values as when I = 0 but fails to keep I constant from setting to setting (see Table 3 for the specific value of I in each one). With the increased value of p→ and the ever-growing barrage of messages that ensues, this independence is now supported by window sizes at least two orders of magnitude lower (though yielding taller peaks) but not significantly by the value of I itself. The value of I, however, does make its influence felt, in the following manner. As mentioned in Section 5, rescaling window size w through a division by T0 makes some of the I > 0 cases stand out for D = 1 by failing to comply (though to a small degree) with the invariance to the values of D and ℓ that seems to be the rule. In that section we correctly attributed this to the increased variance of T0 (through that of ∆(1) ℓ ), but the deviation from invariance clearly increases with the value of I as well (1.81 and 3.06, by Table 3, respectively for settings I and III). On the other hand, the cases for which I < 0 are different in regard to this independence issue. In fact, it is clear from Figure 1 that total-correlation peaks are highly affected by the setting in question. In particular, they are 19 Table 3: Nonzero imbalance values for each of settings I -- IV. Setting I for p→ = 0.01 I for p→ = 0.06 I II III IV −0.53 −0.72 −0.32 −0.78 1.81 0.68 3.06 0.29 increasingly lower than the corresponding peaks for I = 0, and also occur for increasingly lower values of w/T0, as the value of I (see Table 3) becomes more negative. For fixed T0, therefore, this consolidates the value of I as the main driver defining the window duration for which total correlation peaks, provided I ≤ 0. The I = 0 cases are particularly interesting also because total correlation was in all configurations of all settings found to peak roughly between w/T0 = 0.522 and w/T0 = 0.96 (see Table 2). As we noted in Section 5, this is also the range of α, which for I = 0 is the fraction of the reference message output M0T0 yielding the message traffic inside a window of duration w beginning, essentially, at any time t ≥ 0. That total correlation should peak with α between roughly 0.522 and 0.96 seems well aligned with our discussion at the end of Section 3 on maximizing total correlation, through which we found out that a condition necessary (though not sufficient) to such maximization is that only half the nodes receive messages inside the window. Relativizing message traffic inside the window to M0T0 in the I 6= 0 cases must be approached differently, because now fixing α requires the duration of the window starting at time t to either expand with t (if I < 0) or shrink with it (if I > 0). In these cases, taking each w in Table 2 to be the average duration of these windows reveals that total correlation peaks for w/T0 between about 1.47582 and 3.41955 for I < 0 (therefore somewhat above the range recorded for the I = 0 cases), ignoring the more or less degenerate cases of settings II and IV, and between about 0.00101 and 0.00236 for I > 0 (therefore substantially below the I = 0 ranges, as noted above). The corresponding values of α, given by Eq. (27) and illustrated in Figure 4 for setting III, range between 0.71 and 1.917 for I < 0, between 0.007 and 0.017 for I > 0. 6.1 What of the brain? Right at the opening of Section 1 we mentioned the brain as the most represen- tative threshold-based system we know. It is only fitting, then, that we should try and relate it more closely to the model we have developed and analyzed in this paper. This is no simple task for at least two reasons. The first has to do with the model itself, which as we also mentioned in Section 1 has not been meant to faithfully represent the specifics of any one threshold-based system. The second reason is that very little has so far been found out of the brain's so- 20 called microconnectome [36], that is, its network structure at the cellular level, which is where thresholds work their influence. Here we resort to the few glimpses that are available in order to estimate parameter values within our model as well as possible. The largest microcon- nectome to have been mapped to date is from the V1 area of the mouse visual cortex [37]. It reveals a total of 1 278 excitatory neurons and 581 inhibitory neurons, which following our discussion in Section 2.4 allows as to conclude p− ≈ 581/(581 + 1 278) ≈ 0.31, thus agreeing with the commonly accepted range of values for the ratio of inhibitory neurons [38]. It also reveals a total of 29 connected pairs involving 45 neurons, which again in the spirit of Section 2.4 leads to p→ ≈ 29/(45 × 44) ≈ 0.015. Using N = 45 and τ = 7.5 (halfway between 5 and 10, which seem to delimit the range of the typical neuron's ratio of threshold potential to synaptic potential [39 -- 41]), we obtain an imbalance I = −0.96, close therefore to the minimum possible imbalance (I = −1). Note that concluding p→ ≈ 0.015 reflects an effective, rather than merely anatomical, view of the microconnectome. That is, the 29 axon-to-dendrite connections leading to this estimate of the value of p→ do not result simply from the possibility that action potentials travel from a neuron's axon toward one of another's dendrites, but rather from actually observing such potentials.4 This distinction is of paramount importance in the present study, since the parameter p→ is semantically tied to the actual sending of messages in our model (once interpreted as a directed graph, as in Section 2.4), not merely to the existence of a channel through which such messages could be sent. To judge from the values of I for p→ = 0.01 in Table 3, and from the conclusions we have drawn regarding the I < 0 cases of settings I -- IV, this entire attempt to interpret cortical activity in the light of our model seems to be suggesting that we view it as a threshold-based system severely imbalanced on the negative side, therefore incapable of giving rise to any significant amount of total correlation. The missing component, of course, is that the cortex is continually subjected to new input, either external or originating spontaneously from within [45], hence a better characterization of local imbalance in this case would be based on − 1, (32) (N − 1)p→ + µ0 I ′ = µ with µ0 standing for some core amount of external or spontaneous input ex- pected to accompany every µ messages incoming from other cortical neurons. Clearly, µ0 > 0 implies I ′ > I whenever all else remains constant. Local imbal- ance would then be defined through I ′ 6= 0, but without the availability of some 4To further strengthen this notion that the value of p→ is highly network-dependent in terms of effective signaling in the brain, we note that a completely different estimate of p→ can be deduced from a slightly earlier (but similar in the sense of targeting effective connections) study [42], whose authors report a neuron's expected out-degree to be in the order of 100. The study in question is based on the rodent somatosensory cortex, but should a similar conclusion hold for the human cortex taken as a whole, with its 16 billion neurons [43], then an estimate of p→ would place its value at the order of 10−10 and that of I very close indeed to −1. This is all rather speculative, though, especially if we consider the already established fact that different cortical areas of the primate brain have different neuronal densities [44]. 21 rationale on which to base an estimate of µ0 this is essentially as far as we can go. In any event, such redefined imbalance might lie above the I < 0 values of Table 3 while still being negative.5 In this case, we could expect C(X(w)) to peak for w/T0 a few orders of magnitude above 3.41955 (see Table 2). We can go further and estimate the value of T0 as well. Because the cerebral cortex takes up 82% of all brain mass in humans [43], and assuming a uniform density in addition to a brain volume of 1.5 × 10−3 m3 [46, BNID 112053], we can take the cortical volume to be roughly 1.23 × 10−3 m3. In our model, this volume corresponds to the three-dimensional cube of side ℓ ≈ 0.107 m, whence ∆(3) ℓ ≈ 0.071 m (see Section 2.1). Assuming further that action potentials propagate on myelinated axons at a speed (σ) between 50 and 100 m/s [46, BNID 107125] [47] leads, by Eq. (9), to a value of T0 roughly between 0.71 and 1.42 ms. This would place w, the time window duration for peak total correlation, a few orders of magnitude above some value between 2.4 and 4.9 ms. Significant total correlation also arises for w/T0 values below or above the peak's location by a small factor, so window durations of a few hundred milliseconds would support information integration as defined. These, as it turns out, are time lapses within range of the commonly accepted delay before a percept can be rendered conscious [48]. 7 Conclusion Our approach to quantifying the integration of information to which interacting, distributed threshold-based units give rise has relied on a very general model, tailored to no specific system in particular but built around three fundamen- tal notions: that the units interact with one another by passing positively or negatively tagged messages among them; that a unit sends messages out as a function of how the tag balance of incoming messages relates to a threshold; and that this sending out of messages is selective, in the sense of being addressable to a subset of the units only. Assuming that the units are positioned in a one-, two-, or three-dimensional cube uniformly at random, and moreover considering a temporal window of duration w, we have demonstrated by means of extensive computational experiments that information does get integrated in significant amounts inside the window, depending chiefly on the value of w, on the average delay incurred by a message, and on local message imbalance (how much gets transmitted vis-`a-vis what is received, itself dependent on the combined effect of the message-passing and threshold parameters). We have analyzed situations of peak information integration and related the results to cortical dynamics by fixing the model's parameters accordingly. This has served to suggest a vali- dation of the model, despite its purposeful generality, highlighting its potential usefulness for systemic studies of threshold-based interacting units. 5Letting I ∗ denote the greatest of these negative values for I, the purported I ′ such that I ∗ < I ′ < 0 would require 0 < µ0 < µ − (N − 1)p→, assuming for the former of these inequalities that N and all parameters remain unchanged between I ∗ and I ′. 22 Given the window duration w and the set of random variables X(w) (one variable per unit, each related to the reception of messages by the unit in ques- tion within the window), our measure of information integration has been the total correlation of the variables in X(w), here denoted by C(X(w)). This mea- sure seeks to quantify the interdependence of the variables on one another and can be interpreted as information gain that is attained in excess of the total gain the units achieve at the local level. Total correlation can also be interpreted as the Kullback-Leibler (KL) divergence of one probability mass function relative i = xi). That is, to another, viz., of Pr(X(w) = x) relative to QN C(X(w)) can be rewritten as i=1 Pr(X (w) C(X(w)) = Xx∈{0,1}N Pr(X(w) = x) log2 Pr(X(w) = x) i=1 Pr(X (w) QN i = xi) . (33) This expression highlights the well-known fact that the KL divergence is asym- metric with respect to the two mass functions it applies to. Note, however, that this is of no import in our context, because total correlation corresponds to the very specific case in which the KL divergence applies to the two mass functions above and in the direction indicated only. We mention this issue because the current version of the all-partitions theory of integrated information mentioned in Section 1 abandons the KL divergence in favor of the so-called earth-mover's distance because of the latter's symmetry and consequent status as a distance between two probability mass functions [49]. This provides further distinction between the two approaches. We finalize by noting that extensions to our approach are certainly possible, especially by attempting to encompass those systems that, despite embodying explicit references to a threshold-based dynamics, do so in a manner that is not completely aligned with the one our model assumes. One example involves the role of thresholds in the grouping and synchronization of the interacting units [50]. Perhaps our approach can provide useful insight in such contexts as well. Acknowledgments The author acknowledges partial support from CNPq, CAPES, and a FAPERJ BBP grant. References [1] M. F. Bear, B. W. Connors, and M. A. Paradiso. Neuroscience. Wolters Kluwer, Philadelphia, PA, fourth edition, 2015. [2] W. S. McCulloch and W. Pitts. A logical calculus of the ideas immanent in nervous activity. B. Math. Biophys., 5:115 -- 133, 1943. 23 [3] A. Pistorello, C. Romoli, and S. Crespi-Reghizzi. Threshold nets and cell- assemblies. Inform. Control, 49:239 -- 264, 1981. [4] A. Ilachinski. Cellular Automata. World Scientific, Singapore, 2001. [5] J. J. Hopfield. Neural networks and physical systems with emergent col- lective computational abilities. Proc. Natl. Acad. Sci. USA, 79:2554 -- 2558, 1982. [6] J. J. Hopfield. Neurons with graded response have collective computational properties like those of two-state neurons. Proc. Natl. Acad. Sci. USA, 81:3088 -- 3092, 1984. [7] D. H. Ackley, G. E. Hinton, and T. J. Sejnowski. A learning algorithm for Boltzmann machines. Cognitive Sci., 9:147 -- 169, 1985. [8] J. Pearl. Probabilistic Reasoning in Intelligent Systems. Morgan Kaufmann, San Mateo, CA, 1988. [9] T. Hrycej. Gibbs sampling in Bayesian networks. Artif. Intel., 46:351 -- 363, 1990. [10] R. Kinderman and J. L. Snell. Markov Random Fields and Their Applica- tions. American Mathematical Society, Providence, RI, 1980. [11] S. Geman and D. Geman. Stochastic relaxation, Gibbs distributions, and IEEE T. Pattern Anal., 6:721 -- 741, the Bayesian restoration of images. 1984. [12] V. C. Barbosa. Massively Parallel Models of Computation. Ellis Horwood, Chichester, UK, 1993. [13] D. Koller and N. Friedman. Probabilistic Graphical Models. The MIT Press, Cambridge, MA, 2009. [14] P. Smolensky. Information processing in dynamical systems: foundations of harmony theory. In D. E. Rumelhart, J. L. McClelland, and PDP Research Group, editors, Parallel Distributed Processing, volume 1, pages 194 -- 281. The MIT Press, Cambridge, MA, 1986. [15] A. Fischer and C. Igel. Training restricted Boltzmann machines: an intro- duction. Pattern Recogn., 47:25 -- 39, 2014. [16] G. E. Hinton. Deep belief networks. Scholarpedia, 4:5947, 2009. [17] V. Mnih, K. Kavukcuoglu, D. Silver, A. A. Rusu, J. Veness, M. G. Belle- mare, A. Graves, M. Riedmiller, A. K. Fidjeland, G. Ostrovski, S. Petersen, C. Beattie, A. Sadik, I. Antonoglou, H. King, D. Kumaran, D. Wierstra, S. Legg, and D. Hassabis. Human-level control through deep reinforcement learning. Nature, 518:529 -- 533, 2015. 24 [18] Y. LeCun, Y. Bengio, and G. Hinton. Deep learning. Nature, 521:436 -- 444, 2015. [19] D. Silver, A. Huang, C. J. Maddison, A. Guez, L. Sifre, G. van den Driessche, J. Schrittwieser, I. Antonoglou, V. Panneershelvam, M. Lanctot, S. Dieleman, D. Grewe, J. Nham, N. Kalchbrenner, I. Sutskever, T. Lill- icrap, M. Leach, K. Kavukcuoglu, T. Graepel, and D. Hassabis. Master- ing the game of Go with deep neural networks and tree search. Nature, 529:484 -- 489, 2016. [20] A. Nathan and V. C. Barbosa. Network algorithmics and the emergence of information integration in cortical models. Phys. Rev. E, 84:011904, 2011. [21] K. K. Cassiano and V. C. Barbosa. Information integration in elementary cellular automata. J. Cell. Autom., 10:235 -- 260, 2015. [22] S. Watanabe. Information theoretical analysis of multivariate correlation. IBM J. Res. Dev., 4:66 -- 82, 1960. [23] D. W. Zhou, D. D. Mowrey, P. Tang, and Y. Xu. Percolation model of sensory transmission and loss of consciousness under general anesthesia. Phys. Rev. Lett., 115:108103, 2015. [24] D. Balduzzi and G. Tononi. Integrated information in discrete dynami- cal systems: motivation and theoretical framework. PLoS Comput. Biol., 4:e1000091, 2008. [25] J. N. D. Kerr, D. Greenberg, and F. Helmchen. Imaging input and output of neocortical networks in vivo. Proc. Natl. Acad. Sci. USA, 102:14063 -- 14068, 2005. [26] S. Shoham, D. H. O'Connor, and R. Segev. How silent is the brain: is J. Comp. Physiol. A, there a "dark matter" problem in neuroscience? 192:777 -- 784, 2006. [27] A. M. Mathai, P. Moschopoulos, and G. Pederzoli. Random points associ- ated with rectangles. Rend. Circ. Mat. Palermo, 48:163 -- 190, 1999. [28] D. P. Robbins and T. S. Bolis. Average distance between two points in a box. Am. Math. Mon., 85:277 -- 278, 1978. [29] A. Zilinskas. On the distribution of the distance between two points in a cube. Random Oper. Stoch. Equ., 11:21 -- 24, 2003. [30] W. Feller. An Introduction to Probability Theory and Its Applications, volume I. John Wiley & Sons, New York, NY, third edition, 1968. [31] M. Penrose. Random Geometric Graphs. Oxford University Press, Oxford, UK, 2003. 25 [32] P. Erdos and A. R´enyi. On random graphs. Publ. Math. (Debrecen), 6:290 -- 297, 1959. [33] R. M. Karp. The transitive closure of a random digraph. Random Struct. Algor., 1:73 -- 93, 1990. [34] T. S. Han. Multiple mutual informations and multiple interactions in fre- quency data. Inform. Control, 46:26 -- 45, 1980. [35] L. L. Campbell. Averaging entropy. IEEE T. Inform. Theory, 41:338 -- 339, 1995. [36] J. M. L. Budd and Z. Kisv´arday. Communication and wiring in the cortical connectome. Front. Neuroanat., 6:42, 2012. [37] W.-C. A. Lee, V. Bonin, M. Reed, B. J. Graham, G. Hood, K. Glattfelder, and R. C. Reid. Anatomy and function of an excitatory network in the visual cortex. Nature, 532:370 -- 374, 2016. [38] B. Rudy, G. Fishell, S. Lee, and J. Hjerling-Leffler. Three groups of in- terneurons account for nearly 100% of neocortical GABAergic neurons. Dev. Neurobiol., 71:45 -- 61, 2011. [39] M. Steriade, I. Timofeev, and F. Grenier. Natural waking and sleep states: a view from inside neocortical neurons. J. Neurophysiol., 85:1969 -- 1985, 2001. [40] A. Hasenstaub, Y. Shu, B. Haider, U. Kraushaar, A. Duque, and D. A. Mc- Cormick. Inhibitory postsynaptic potentials carry synchronized frequency information in active cortical networks. Neuron, 47:423 -- 435, 2005. [41] Y. Shu, A. Hasenstaub, A. Duque, Y. Yu, and D. A. McCormick. Modula- tion of intracortical synaptic potentials by presynaptic somatic membrane potential. Nature, 441:761 -- 765, 2006. [42] M. Shimono and J. M. Beggs. Functional clusters, hubs, and communities in the cortical microconnectome. Cereb. Cortex, 25:3743 -- 3757, 2015. [43] S. Herculano-Houzel. The human brain in numbers: a linearly scaled-up primate brain. Front. Hum. Neurosci., 3:31, 2009. [44] C. E. Collins, D. C. Airey, N. A. Young, D. B. Leitch, and J. H. Kaas. Neuron densities vary across and within cortical areas in primates. Proc. Natl. Acad. Sci. USA, 107:15927 -- 15932, 2010. [45] A. Luczak, P. Bartho, and K. D. Harris. Gating of sensory input by spon- taneous cortical activity. J. Neurosci., 33:1684 -- 1695, 2013. [46] R. Milo, P. Jorgensen, U. Moran, G. Weber, and M. Springer. BioNumbers -- the database of key numbers in molecular and cell biology. Nucl. Acids Res., 38:D750 -- D753, 2010. 26 [47] H. A. Swadlow and S. G. Waxman. Axonal conduction delays. Scholarpedia, 7:1451, 2012. [48] M. H. Herzog, T. Kammer, and F. Scharnowski. Time slices: what is the duration of a percept? PLoS Biol., 14:e1002433, 2016. [49] M. Oizumi, L. Albantakis, and G. Tononi. From the phenomenology to the mechanisms of consciousness: integrated information theory 3.0. PLoS Comput. Biol., 10:e1003588, 2014. [50] K. P. O'Keeffe, P. L. Krapivsky, and S. H. Strogatz. Synchronization as aggregation: cluster kinetics of pulse-coupled oscillators. Phys. Rev. Lett., 115:064101, 2015. 27
1807.10374
1
1807
2018-07-26T21:37:02
Towards Identifying the Systems-Level Primitives of Cortex by In-Circuit Testing
[ "q-bio.NC" ]
The hypothesis considered here is that cognition is based on a small set of systems-level computational primitives that are defined at a level higher than single neurons. It is pointed out that for one such set of primitives, whose quantitative effectiveness has been demonstrated by analysis and computer simulation, emerging technologies for stimulation and recording are making it possible to test directly whether cortex is capable of performing them.
q-bio.NC
q-bio
Towards Identifying the Systems-Level Primitives of Cortex by In-Circuit Testing Leslie G. Valiant [email protected] Harvard University July 22, 2018 Summary The hypothesis considered here is that cognition is based on a small set of systems-level computational primitives that are defined at a level higher than single neurons. It is pointed out that for one such set of primitives, whose quantitative effectiveness has been demonstrated by analysis and computer simulation, emerging technologies for stimulation and recording are making it possible to test directly whether cortex is capable of performing them. 1. Introduction In both biology and computer science the virtue of modularity in complex systems is widely understood. In contrast, the idea that complex systems need to have many levels of description, while well appreciated in computer science, has been much less discussed in biology. We believe that for a computational theory of cortex to be useful, it would need to give an account of algorithmic tasks at levels intermediate between those of neurons and behavior. Just as we cannot understand the workings of a cell phone by studying its operations only at the single bit level, we cannot expect to understand the brain at the single neuron level. The suggestion we explore here is that cognitive computation in cortex is built on a small collection of base primitives at the systems level. This contrasts with Hebbian plasticity (Hebb, 1949), which is a single primitive at the single neuron level. Here we discuss one candidate for such a collection of base primitives, one distinguished by the existence of demonstrations, by analysis and computer simulations, that it offers a scheme in which the capacity of cortex for basic cognitive tasks can be given a quantitative explanation (Valiant 1994, 2005, 2006; Feldman & Valiant 2009). This collection consists of: association, memorization, inductive learning, and hierarchical memory assignment, all suitably defined. The purpose of this note is to point out that because of technological advances over the last decades, especially in optogenetics (Miyawaki et al. 1997, Boyden et al. 2005, Deisseroth et al. 2006, Shemesh et al. 2017), it is now becoming feasible to test systematically whether cortex can indeed perform these primitives. 1 These four primitives are defined below in terms of their actions on randsets, random sets of neurons of a certain size. The nature of the proposed experiments is analogous to "in-circuit testing" of a traditional electronic printed circuit board, where instead of testing the input-output behavior of the whole board, the testing is done on smaller parts, using only probes. In the proposed experiments there will be no sensory stimuli or behavioral responses -- the experiments will consist entirely of stimulating and recording from appropriate sets of neurons. The goal is to find out what changes in computational function in a subcircuit of cortex can be achieved by specific neural stimulation. Existing experimental evidence that the firing patterns of cortical neurons can be altered by suitable behavioral training (Jackson et al. 2006, Rebesco et al. 2010) is relevant and encouraging, but does not directly address the question of which computational primitives are being executed. Behaviors prompted by sensory inputs cause all kinds of activity in the brain extraneous to the execution of any one primitive. The in-circuit methodology is suggested to isolate the operation of a subcircuit as much as possible from such interference. The goal therefore is to test, by direct neural stimulation and recording, whether cortex is capable of performing certain basic computational tasks. The paradigmatic first such task is "association of a set A to set B", abbreviated as A→B, which here is defined as follows: for neuron sets A and B we want to induce the new functionality that, after the training, whenever set A is stimulated then the newly modified circuit will cause set B to become active also. Training here would involve stimulating A and B according to some timed protocol. Testing would involve stimulating A again and now recording from the set B that had been stimulated in training. This particular task may be viewed as a direct systems level analog of Hebbian plasticity. We note that our notion of randsets is not the same as the notion of assemblies, in the sense of Hebb (1949). That traditional notion of assemblies is associated with the idea that the neurons in these assemblies are better connected to each other than to other neurons (e.g. Carillo- Read et al. 2017). Randsets account for the claimed computational capabilities simply by virtue of being randomly chosen, and probably with weak criteria of randomness being sufficient. There is no requirement that they be well connected to each other. The theory basically asserts that for random sets in a randomly connected network there will be a large enough bandwidth of reliable communication from one randset to another, that simple local updates at the neurons that result from training will be sufficient to realize the claimed functionalities. There is no comparable theory that shows that the high interconnectivity assumption of traditional assemblies provides any comparable computational capabilities. Fortunately, also, it is the randset theory that appears to be one that is readily testable experimentally for the sets of operations we contemplate, since random sets are more easily identified than highly interconnected ones. (Of course, it may be that our randsets, once selected, do become more strongly connected to each other over time, but that is not our concern here.) An experiment relevant to what we are contemplating here is reported in Seeman et al. (2017), where the setting up of such associations was attempted by electronic stimulation, but did not succeed reliably. The experiment was done for one value of r. Possible interpretations are that 2 the value of r used there was too small, or that it was not a part of cortex that performs this function. We are suggesting a systematic exploration of the relevant parameter space, and in particular the set size r. Another relevant result, one obtained by optogenetic means, is by Carrillo-Reid et al. (2016), which can be viewed as also an example of "in-circuit" testing, but the task considered there was pattern completion, as in auto-associative nets (e.g. Hopfield, 1982). We believe that pattern completion is a less fundamental operation for our purposes than the ones we consider here, since there is no clear definition of what it should be, and any reasonable system offers some version of it as an epiphenomenon. 2. The Four Tasks and their Potential Validation Each instance of the tasks described here can be induced by an associated training protocol, and is realized as local changes in the neurons within the given randomly connected network, giving rise to a subcircuit having the functionality of that task instance. Computer simulations (Feldman & Valiant, 2009) show that thousands of instances of such tasks can be realized with realistic network parameters without having the continued effectiveness of the earlier ones acquired degraded by later ones. We note that having multiple task types, which between them may be adequate to form a basis for cognition, is a much more exacting requirement than single-task models such as traditional "auto-associative nets" which perform the single task of pattern completion. The longer term motivation is that of identifying a "cognitively adequate'' set of primitives, perhaps our four forming the core, and showing that higher level cognitive operations can be efficiently implemented in terms of these. The following detail the four tasks that we are suggesting as the core, and for which we are here suggesting validation by in-circuit testing. An example of some more detailed specifications still, for one incarnation, is described by Feldman and Valiant (2009). Association: For sets A, B train so that result is: If in future A is stimulated then B (but not random other neurons) will become active. Supervised Memorization of Conjunctions: For sets A, B, C train so that the result is: If in future A and B are stimulated then C will become active also (but not if just one of A, B is stimulated.) Inductive Learning of Simple Threshold Functions: For sets A, B, C, D (say) train so that the resulting circuit generalizes a classification at target A, beyond the examples seen, according to a linear separator on the variables B, C, D. In particular, if in future some subset of B, C, D is stimulated then the target A will become active according to whether some linear separator consistent with the examples holds. For example such a linear separator may be B + C + 2D ≥ 2, where now these variables have a 1/0 interpretation corresponding to whether the randsets are active or not. 3 Hierarchical Memory Formation: For sets A, B, train so that result is: a set C (not specified by the experimenter) has been allocated by the computation internally so that the effect "supervised memorization of conjunctions" is achieved for A, B and that, now unknown, C. This realizes the fundamental psychological task of "chunking", where a new compound concept, here A & B, becomes equal citizen in the circuit with earlier ones. While the first three of these tasks can be tested directly by in-circuit testing, the last task appears more difficult. One possible approach is to glean information using the timing mechanism at work. Verifying how these tasks are realized in cortex, if indeed they are, is a further area of challenge. The suggestion has been made (Valiant, 2012) that the hippocampal system is responsible for determining the identity of the randset C in hierarchical memory formation, for given A and B. Such a statement, which assigns responsibility for performing a particular task to a particular brain area, may be easier to test. 3. General Considerations How large sets? In the analysis the parameter r is dependent on the number of neurons in the system, the number of connections, the synaptic strengths, and the algorithms used for realizing the tasks. For the association task, for example, having r too small will fail to train the target B as desired, while having r too large will train too many spurious neurons in addition. Hence, we expect that considerable experimentation may be needed to find the right value of r in any one cortical area. Note that in this case of associations, the size of the source set A appears to be critical, but this is the easier set in that it only needs stimulation. The size of the target set B, which has to be both stimulated and recorded from, may be much less critical. Which brain areas? To test our theory one needs to perform the experiment in a cortical area where the corresponding task is implemented in biology. We have behaving mammals in mind. It may be that primates are the best at some of these tasks. There is some uncertainty as to which parts of cortex perform these various tasks, and some experimentation there may be necessary. How sets chosen? Since the hypothesis requires that the experiment work for randomly chosen sets of a certain size in the appropriate brain areas, it is reasonable to go with whatever bias the experimental technique used imposes. Neurons in particular layers may need to be found. The theory aims to explain how large numbers of task instances can be processed. Hence, experiments need to be done in parts of cortex that are large enough that it is plausible that so many tasks are indeed supported. The various sets A, B, C, etc., may be in the same cortical area, or in different areas. What timing protocols for training? Timing may be critical. When stimulating a pair of sets A, B for association we need an asymmetric protocol, such as B being stimulated with one millisecond delay after A. 4 Comparisons? Comparative studies for different cortical areas and different species, including humans, would be of great interest. 4. Conclusion The hypothesis to be tested is that cortex is able to perform a specific set of tasks on random sets of neurons. The new opportunity is offered by the remarkable recent advances in stimulation and recording technologies. The experiments would determine whether these tasks can be performed in appropriate cortices, but not how. However, positive results would demonstrate that the brain has impressive computational capabilities that are relevant to cognitive computation. These systems- level capabilities, if present, are unlikely to have arisen by accident. It would be a remarkable coincidence if the brain had these capabilities but did not use them. 5. Acknowledgement This work was supported in part by a grant from the National Science Foundation CCF 1509178. 6. References Hebb, D. O. (1949). The Organization of Behavior: A Neuropsychological Theory. New York: Wiley and Sons. Valiant, L.G. (1994). Circuits of the Mind, Oxford University Press: New York. Valiant, L.G. (2005). Memorization and association on a realistic neural model. Neural Computation, 17:3, 527-555. Valiant, L. G. (2006). A quantitative theory of neural computation. Biological Cybernetics, 95:3 205-211 Feldman, V. & Valiant, L. G. (2009). Experience-induced neural circuits that achieve high capacity. Neural Computation, 21:10, 2715-54. Miyawaki, A., Llopis, J., Heim, R., McCaffery, J. M., Adams, J. A., Ikura, M., and Tsien, R. Y. (1997). Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin. Nature. 388 (6645): 882 -- 887. Boyden, E. S.; Zhang, F.; Bamberg, E.; Nagel, G.; Deisseroth, K. (2005). Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci. 8 (9): 1263 -- 8. 5 Deisseroth, K., Feng, G., Majewska, A. K., Miesenbock, G., Ting, A., and Schnitzer, M. J. (2006). Next-Generation Optical Technologies for Illuminating Genetically Targeted Brain Circuits. Journal of Neuroscience. 26 (41): 10380 -- 6. Shemesh, Or A, Tanese, D., Zampini, V., Linghu, C., Piatkevich, K., Ronzitti, E., Papagiakoumou, E., Boyden, E.S., and Emiliani, V. (2017). Temporally precise single-cell resolution optogenetics, Nature Neuroscience 20:1796 -- 1806. Jackson, A., Mavoori, J. and Fetz, E. E., (2006) Long-term motor cortex plasticity induced by an electronic neural implant. Nature 444, 56 -- 60. Rebesco, J. M., Stevenson, I. H., Körding, K. P., Solla, S. A., and Miller, L. E., (2010) Rewiring neural interactions by micro-stimulation, Frontiers in Systems Neuroscience, 4:39, pp. 1-15. Carillo-Reid L., Yang, W., Miller, J. K., Peterka, D. S., and Yuste R. (2017) Imaging and Optically Manipulating Neuronal Ensembles, Annual Review in Biophysics, 46:271-293. Jackson, A., Mavoori, J. and Fetz, E. E., (2006) Long-term motor cortex plasticity induced by an electronic neural implant. Nature 444, 56 -- 60. Rebesco, J. M., Stevenson, I. H., Körding, K. P., Solla, S. A., and Miller, L. E., (2010) Rewiring neural interactions by micro-stimulation, Frontiers in Systems Neuroscience, 4:39, pp. 1-15. Seeman, S.C., Mogen, B.J., Fetz, E.E., and Perlmutter, S.I. (2017) Paired Stimulation for Spike- Timing-Dependent Plasticity in Primate Sensorimotor Cortex, Journal of Neuroscience, 37(7):1935 -- 1949. Carillo-Reid, L., Yang, W., Bando, Y., Peterka, D. S., and Yuste, R. (2016) Imprinting and recalling cortical ensembles, Science 353:6300, 691-694. Hopfield, J. J., (1982) Neural networks and physical systems with emergent collective computational abilities", Proceedings of the National Academy of Sciences of the USA, 9:8, 2554 -- 2558. Valiant L. G. (2012) The hippocampus as a stable memory allocator for cortex, Neural Computation, 24:11, 2873-2899. 6
1607.05530
2
1607
2016-09-15T21:58:12
Verbal Perception and the Word Length Effect
[ "q-bio.NC" ]
A theoretical framework is proposed for the understanding of verbal perception -- the conversion of words into meaning, modeled as a compromise between lexical demands and contextual constraints -- and the theory is tested against experiments on short-term memory. The observation that lists of short words are recalled better than lists of long ones has been a long-standing subject of controversy, further complicated by the apparent inversion of the effect for mixed lists. In the framework here proposed, these behaviors emerge as an effect of the different level of localization of short and long words in semantic space. Events corresponding to the recognition of a nonlocal word have a clustering property in phase space, which facilitates associative retrieval. The standard word-length effect arises directly from this property, and the inverse effect from its breakdown. An analysis of data from the PEERS experiments (Healey and Kahana, 2016) confirms the main predictions of the theory. Further predictions are listed and new experiments are proposed. Finally, an interpretation of the above results is presented.
q-bio.NC
q-bio
Verbal Perception and the Word Length Effect Francesco Fumarola (Dated: October 6, 2018) A theoretical framework is proposed for the understanding of verbal perception -- the conversion of words into meaning, modeled as a compromise between lexical demands and contextual constraints -- and the theory is tested against experiments on short-term memory. The observation that lists of short words are recalled better than lists of long ones has been a long-standing subject of controversy, further com- plicated by the apparent inversion of the effect for mixed lists. In the framework here proposed, these behaviors emerge as an effect of the different level of localization of short and long words in semantic space. Events corresponding to the recognition of a nonlocal word have a clustering property in phase space, which facilitates associative retrieval. The standard word-length effect arises directly from this property, and the inverse effect from its breakdown. An analysis of data from the PEERS experiments (Healey and Kahana, 2016) confirms the main predictions of the theory. Further predictions are listed and new experiments are proposed. Finally, an interpretation of the above results is presented. I. INTRODUCTION A. Free recall: lexical and serial-position effects In 1894, with their pioneering work on free-recall experiments, Binet and Henry intro- duced a key tool for the controlled investigation of short-term memory (Binet and Henry, 1894). In its traditional form, a free-recall experiment is performed by presenting the subject with a list of words and then requesting him or her to recall it in any order (Murdock, 1960, 1962; Roberts, 1972; Standing, 1973). Several types of effects have been reported: 1) Effects depending on the lexical properties of individual words. In particular, lists of short words are recalled better than lists of long ones, a fact known in the literature as the word-length effect (Baddeley et al., 1975; Russo and Grammatopoulou, 2003; Tehan and 6 1 0 2 p e S 5 1 ] . C N o i b - q [ 2 v 0 3 5 5 0 . 7 0 6 1 : v i X r a 2 Tolan, 2007; Bhatarah et al., 2009). 2) Effects in which the recall probability depends on the absolute position of words in the list. It has been observed that the first and last words in the list are recalled more easily ("primacy" and "recency" effects). 3) Effects depending on the relative position of words with respect to each other. Most notably, the recall probabilities of contiguous words correlate positively, a fact known as the contiguity effect (Murdock, 1960, 1962). The need to understand serial-position effects led to the devising of retrieved-context models, such as the Temporal Context Model (Howard and Kanaha, 2002). In these models the recall process, rather than retrieving a word directly, retrieves the context associated to the word first. Within this scenario, recency effects appear because the context at the time of the "mem- ory test" is most similar to the context associated with recent items. When an item is re- trieved, it reinstates the context active when that item was presented. Because this context overlaps with the encoding context of the items' neighbors, a contiguity effect results. Through these models, serial-position effects have been substantially understood over the past fifteen years (Howard and Kahana, 2002, 2002b; Sederberg, Howard, and Kahana, 2008; Polyn, Norman, and Kahana, 2009, 2009b; Lohnas, Polyn, and Kahana, 2015; Kahana, 2012). The same cannot be said, however, about the word-length effect. B. Riddles of the word length effect The word-length effect (WLE) has been a traditional testing ground for models of short- term memory (Campoy, 2011; Jalbert et al., 2011), and it has played a key role in establishing the working-memory paradigm and the phonological loop hypothesis (Baddeley and Hitch., 1974). The standard account of the effect (Baddeley, 2007) relies on a trade-off between memory decay (in the phonological store) and subvocal rehearsal via an articulatory control process. Because shorter words take less time to rehearse, more decaying traces of them can be refreshed than decaying traces of long items, and, therefore, more short items can be recalled. This picture, however, is not able to account for all experimental observations concerning this effect, and has been repeatedly called into question. 3 In (Neath et al., 2003), it was shown that with words having the same number of syllables but different pronunciation times, no unambiguous WLE arises. This result (extended in Jalbert et al. 2011) suggests that the effect depends on the number of syllables, and not on the time it takes to pronounce them. Experiments have also been performed in conditions where there was a delay between lists, making subvocal rehearsal possible in the interval. No appreciable difference in recall probabilities was found (Campoy, 2008). In the same study, experiments were performed in which subvocal rehearsal was prevented by a high presentation rate. No delay was allowed between the presentation of word lists and the memory test. Yet, the WLE occurred unperturbed. In the 2011 paper I just cited, Jalbert et al. concluded: "the WLE may be better explained by the differences in linguistic and lexical properties of short and long words rather than by length per se". C. Semantic predictors of word length In the meanwhile, within the fields of experimental and computational linguistics, progress has been made in understanding the role of word length in verbal processing. Over the years, it has emerged that words with different lengths tend to have different semantic properties. The idea was first put forth in pedagogical studies (Klare, 1988). In (Elts, 1995), a correlation coefficient of 0.96 was found between a noun's length and its average tendency to be used as a technical term ("terminologicality"). Mikk et al. (2000), using data on the human-assessed complexity of a large sample of words, found a correlation coefficient 0.86 between words' length and their semantic complexity. Pinning down the precise semantic property that correlates to word length has proven difficult. Already in (Greenberg, 1966) it was argued that a word's length correlates pos- itively to its conceptual "markedness" of meaning. Various notions of markedness have subsequently been discussed in the literature (Haspelmath, 2006). Piantadosi et al. (Piantadosi et al., 2011, 2011B) and later Mahowald et al., (Mahowald et al., 2012) reported that the length of words correlates positively with their contextual information rate. More recently, Lewis and Frank (2016) have carried out a comprehensive 4 experimental study across 80 languages. They found that, in all the languages considered, judgments of conceptual complexity for a sample of real words correlate highly with their length, and they even control for frequency, familiarity, imageability, and concreteness. Their conclusion is: "While word lengths are systematically related to usage − both frequency and contextual predictability − our results reveal a systematic relationship with meaning as well". In the light of these findings, it would be a natural step to attempt an explanation of the WLE in terms of the semantic differences among words. However, no such approach seems to have been attempted in the literature. D. The inverse word length effect Recently, new aspects of the WLE have emerged through the analysis of a large set of data from experiments by Miller and al (Miller et al., 2012). The data analysis was performed by Katkov et al. (Katkov et al., 2014), who found no negative correlation between total length of presented items and number of recalled words, thus disproving both rehearsal-time theories and hypotheses based on the increasing complexity of longer items. Moreover, they reported an inversion of the effect in mixed lists, that is, lists where words are selected irrespectively of their length. They observed that, in this type of lists, the mean values of recall probabilities allow to establish an increasing trend. Long words are recalled better than short ones. An "inverse" WLE had been previously reported by at least two groups, but in somewhat less general circumstances: one of them (Hulme et al., 2006) embedded strictly pure lists with a single word of a different type, while the results of Xu et al. (Xu et al., 2009), may not bear direct comparison with data in languages other than Chinese. If the inversion of the WLE for mixed lists will be confirmed by further experiments, it will have to be taken into account by every general theory of the standard WLE. Let us consider, therefore, what requirements a model should fulfill to explain both phenomena simultaneously. Call γ the fraction of long words in the list; call Pl(γ) the probability of recalling suc- cessfully a given long word from a list in which a fraction γ of words are long; and let Ps(γ) be the probability of recalling successfully a short word, from a list with a fraction γ of long 5 words. Obviously, the function Pl(γ) is only defined for γ > 0, and the function Ps(γ) only for γ < 1. For γ ∈]0, 1[, both functions are defined. Theorists would have to reconcile two observations on the curves Ps(γ), Pl(γ): 1. Pl(γ = 1) < Ps(γ = 0) 2. Pl(γ) > Ps(γ) for γ ∈]0, 1[. The only way these two inequalities can be simultaneously satisfied is if both Pl and Ps are, on the whole, decreasing functions of γ. The simplest choice of these curves compatible with experiments is one where both are monotonously decreasing, that is: dPl dγ < 0 dPs dγ < 0 (1) Pl Ps γ = 0 γ = 1 FIG. 1: Back-of-envelope sketch of the possible behavior of recall probabilities for short and long words, as a function of the fraction of long words in the list. This means that, whenever we replace a short word of the list with a long word, we are lowering the recall probability of all the words in the list, both long and short. A higher number of long words makes every single word in the list harder to recall. It is difficult to imagine how this could ensue from the different duration of long and short words. The question is then: can equation (1) result directly from the different semantic properties of long and short words? 6 In this paper, I will show that the answer is positive as long as one models carefully the process of verbal perception. A suitable way of doing so is demonstrated in the next section. In section III, I employ a retrieved-context description of verbal recall to derive both WLEs (standard and inverse). In section IV, I test two key predictions of the theory against data from the PEERS experiment of Kahana et al. (Lohnas and Kahana, 2013; Healey and Kahana, 2016). In section V, I list five experiments designed to test further predictions of the theory. In the Conclusions, I sketch a possible interpretation of the results. II. VERBAL PERCEPTION A. Structure of Semantic Space Verbal perception is the conversion of words into meaning, and any theory of the phe- nomenon must begin by defining the space in which the mental trajectory takes place. I will refer to this as "semantic space" and will represent "semantic states" as integers belonging to a segment X = (−A, A), with A ∈ N. Measurable quantities are to be computed in the limit A → ∞, and a generalization to higher dimensions will be introduced below. Since X ⊆ Z, it follows that, between two discrete times t0 and t0 + N , the psychological trajectory can be represented by a sequence of integers x0, x1, . . . , xN . As memory plays a crucial role in thought process, the laws of motion governing the trajectory will be in general highly non-markovian. Call W the vocabulary available to the system. Naturally, a given word w ∈ W may be appropriate to describing more than one state, though with a varying degree of appropri- ateness. For each word w, thus, one can define its semantic range Xw ⊂ X as the set of all states described by that word. Meanings should be seen as distributed in an universal way over X, as in Figure 2. The function w → Xw matches each word with locations associated to meanings suitable for that word. Not all states are equally fit to be verbalizable. Hence, there will be a varying probability qx that the state x will match the word describing it. The value of qx is the degree of verbalizability of state x; or conversely, given x ∈ Xw, qx is the fitness of word w as a descriptor of state x. 7 FIG. 2: Position dependence of word distribution in semantic space. The space is represented as an array of boxes. Each box represents a state, each color a word. Three of the words are typed above next to circles the color that represents them. The images underneath show meaning in different regions. States are filled with color to the extent to which they are verbalizable. The white boxes correspond to qx = 0. (cid:83) If the sets {Xw}w∈W do not overlap, we can define a verbalization function vx : w∈W Xw → W such that vx = w whenever x ∈ Xw. When the system tries to ver- balize state x ∈ X, it produces word vx with probability qx, and the silent word 0 with probability 1 − qx. We will define qx so that it is only null for nonverbalizable states. Finally, we can take the trajectory to become markovian whenever it is driven by a specific verbal input: at the instant in which a new word is presented, the system hops into the nearest state described by that word. Accordingly, let the function ξw : X → Xw be defined by the condition: (2) with the understanding that, if two x ∈ Xw satisfy this condition, ξw(x) is chosen by ξw(x) − x ≤ x − x ∀x ∈ Xw flipping a coin. Given a verbal input (cid:126)w = (w1, w2, . . . , wN ), with the wi's ∈ W, and a starting point x0, the trajectory of the system while perceiving the input will be given by : = "tail"= "tusk"= "neck" 8 FIG. 3: Representation of the trajectory induced by the perception of a verbal input. The input consists of five words, depicted here as circles. As the system perceives each new word, it reaches for the nearest state described by that word. xn = ξwn ◦ ξwn−1 ◦ . . . ◦ ξw1(x0) (3) ∀n ∈ (1, N ) . Obviously, the ordering of words in the input is essential. Suppose for instance that the vocabulary contains two words, W = {A, B}, with XA = {0, 3} and XB = {2}, and the starting point is x0 = 0. The verbal input (A, B) yields the trajectory (0, 0, 2), while the input (B, A) yields the trajectory (0, 2, 3). Word A corresponds to two different states in the two inputs. FIG. 4: Noncommutative nature of verbal perception. Example of trajectories induced by the permutation of two words. B. Effect of the Voronoi length distribution The verbalizable states associated to word w can be seen as seeds of a Voronoi partition of X into cells {cx}x∈Xw, where states located on the boundaries are understood to belong to each of two cells with probability 1 2. 9 FIG. 5: Voronoi cells of the "red" and "green" word within the same system. During the perception of verbal inputs, the distance travelled by the system is controlled by the function d(x, w) := ξw(x)− x, the "reaching distance" of word w from state x. This is just the distance of x from the closest state belonging to Xw. The level of nonlocality of a word w may be measured by its reaching distance averaged over all starting points: dw = 1 X (cid:88) x∈X ξw(x) − x (4) Obviously dw1 > dw2 does not necessarily imply Xw1 > Xw2. Consider two words w1, w2 such that Xw1 = {−A, A} and Xw2 = {−A/2, A/2}, with A even and > 2. We have Xw1 = Xw2, but dw1 = A2 2A+1 > dw2 = A2+A 2(2A+1). A direct relation can be shown to exist between a word's average reaching distance and the size of its Voronoi cells. Call λi the size of the i-th Voronoi cell of word w. Assuming ¯λ (cid:29) 1, we may neglect boundary effects and write (cid:88) i∈Xw λi/2(cid:88) d=1 2 dw ∼ 1 X (cid:32) d ∼ ¯λ 4 (λ − ¯λ)2 ¯λ2 1 + (cid:33) (5) Thus, the average reaching distance of a word depends solely on the first two moments of its Voronoi length distribution. If (λ − ¯λ)2 (cid:38) ¯λ2 for a word w, its word structure contains strong fluctuations, so one may separate X into regions were states described by w are denser, and regions were they are sparser. Word w is effectively "localized" inside the regions were such states are dense, that is, it expresses those semantic areas better than those where its states are sparse. The degree of word localization has arguably a strong effect on the dynamics, as shown in Figure 6. 10 FIG. 6: Example of two trajectories induced by the same input (shown as an array of circles). The two trajectories begin when the system in different states. The red (localized) word creates a narrow trapping region in semantic space, marked by the dashed ellipse. Once the localized word appears in the input, all trajectories are trapped within the ellipse. C. Analysis of word-length related properties In most experiments, the relevant number of word-lengths is four, since words with more than four syllables are rare in English. For simplicity, here we will consider the existence of only two word-lengths, short and long. Call dα (with α = s, l) the reaching distance averaged over space and over all words of the same length (short or long). From eq. (5), we see that the space-averaged reaching distance is the product of two factors, involving respectively the first and second moment of the Voronoi length distribution. Let us consider how these two moments depend on the word's length. The average Voronoi length ¯λ may be related to the frequency of a given word in a corpus of the language. Indeed, if the system, in its 'speaking mode', explores semantic space ergodically and uniformly, the frequency of a word is νw = λw −1. In a typical corpus of the English language, the frequency ν(S) of words with S syllable is monotonously decreasing. As a consequence, we expect the average of the Voronoi length λw to be larger for short words than for long words. While this is correct for most languages, notice that there are exceptions, such as Turkish and Arabic, where the function ν(S) is peaked at S = 2 (Fucks, 1956; Grzybek, 2007). 11 Let us now look at the relative fluctuations of the Voronoi length, described by (λ−¯λ)2 2 λ . We will surmise their magnitude through a qualitative argument. As mentioned in the Introduction, various approaches have been taken to prove that long words are on average more 'technical', 'specialized', 'distinctive' or 'marked' than short words. Several claims made in (Elts, 1995), (Mikk et al., 2000), and (Lewis and Frank, 2016) may be rephrased as the statement that long words are, on average, conceptually more specific. A word is conceptually specific if it is localized in certain areas of semantic space. A correlation exists, therefore, between word-length and semantic localization. Localization, in turn, will occur if the scale over which the Voronoi length fluctuates is comparable or greater than its average value. This leads to the conclusion that the relative fluctuations of the Voronoi length will be larger for longer words. Thus, both factors in eq. (5) take a greater value if the word is long. It follows that dl > ds. D. Higher dimensions A one-dimensional modeling of semantic space is of course unrealistic, and may not suffice for every application of the theory. If X is taken to be a connected subset of Zn, the definition we have given for the word structure {Xw}w∈W applies all the same. The points in space are now vectors, and equations (3) and (4) are still valid, the distance in eq. (4) being the Euclidean distance in n dimensions. The Voronoi cells, however, become less simple to treat as they can be arbitrary polyhedra (for a complete treatement, see Aurenhammer et al., 2013). Formula (5) for the reaching distance must be modified, and it takes a geometry-dependent form. The Voronoi structure, yet, affects directly only the process of verbal perception, not the process of memory retrieval, which will be the subject of the next section. Thus, while in the figures I will refer to the one-dimensional case, the mathematical results will apply to any number of dimensions. 12 III. VERBAL RECALL A. Retrieved-context primer We have seen that the rules of motion of the system become markovian during the perception of verbal input. In retrieved-context models of short-term memory, the rules of motion are also markovian during the search for memories, which is conducted through the principle of free association. In these models, the retrieval of memories per se is not a measurable phenomenon. What can be observed is the retrieval of words describing those memories. Therefore, a recall experiment must be seen as the composition of two processes: a processes of memory retrieval, and a process of memory verbalization. The markovian process of memory retrieval must be described here, in keeping with the spirit of context-driven models, as a random walk on X that effects a retrieval whenever it meets a state corresponding to the experience to be recalled. The verbalization process, on the other hand, depends on the verbalizability of memories: a memory x, once retrieved, has a probability qx of leading the system to produce the word describing it. The following mathematical problem arises. Supposing one is given 1) the structure {vx, qx}x of the vocabulary; 2) a word list (cid:126)w = (w1, w2, . . . , wN ) presented to the system; 3) the state of the system when the word list begins to be presented, that is, a probability distribution χ[y0] on its position y0; 4) the state of the system when the retrieval process begins, that is, a probability distri- bution ψ[x0] on the new position x0; one wants to predict the probability that the i-th word will be among those recalled by the system. In the next section, this program will be carried out for the particular case of a vocabulary containing words of two different lengths. B. Application to the double word-length scenario We will begin by defining the probability Ph(t) that a memory placed at distance h from x0 is met by the retrieving random walk for the first time after a time t. In one dimension, this is given by 13 y0 x0 FIG. 7: Trajectory during presentation and recall of a three-word list. Starting from a random position (stage A), the system moves under the effect of verbal input (stage B), and its discontinuous path leaves memory traces (stage C) that are pursued by a random walk in the retrieval stage (stage D). The points in space-time corresponding to retrieval are boxed. (cid:88) (cid:126)n: (cid:80)h 1 ni=t Ph(t) = fn1fn2...fnh where f2n = 0 and f2n−1 = (2n−3)!! n!2n . ABDtC The probability that a memory will be retrieved is therefore T(cid:88) pretrieval(h) = Ph(t) 14 (6) where the cutoff T is needed to obtain meaningful results in one or two dimensions, and 0 can otherwise be let to infinity. Suppose a list (cid:126)w = (w1, w2, . . . , wN ) has been presented. Call pi( (cid:126)w; y0, x0) the probability of retrieving the memory created by the i-th word in the list, given a certain initial position y0 for the trajectory during presentation, and a certain initial position x0 for the trajectory during retrieval. We have: (cid:18)(cid:12)(cid:12)(cid:12) Ξi(y0) − x0 (cid:12)(cid:12)(cid:12)(cid:19) pi( (cid:126)w; y0, x0) = pretrieval (7) where Ξi := ξwi ◦ . . . ◦ ξw2 ◦ ξw1. Our goal is to compute the average recall probability for an arbitrary list composed by S short words and L long words arranged into a given order. This amounts to averaging eq. (7) over all lists of the same type (cid:126)α = (α1, . . . , αN ) where αi ∈ {s, l} for i = 1, . . . , N : pi((cid:126)α; y0, x0) = 1 s W L W S l pi( (cid:126)w; y0, x0) := pretrieval (cid:42) (cid:18)(cid:12)(cid:12)(cid:12) Ξi(y0) − x0 (cid:12)(cid:12)(cid:12)(cid:19)(cid:43) (8) (cid:126)α (cid:88) (cid:126)w∈W N α(wi)=αi i=1,...,N where α(w) is the type of word w and Ws (Wl) is the number of short (long) words in the vocabulary. We will perform the averaging over lists through a mean-field approach. Mean-field approaches, widely employed in physics, consist in inverting the order of the two steps: the computation of observables and the averaging. Instead of averaging the final probabilities, one averages an intermediate, non-observable quantity usually called the "field". In this case, we may average the functions ξw themselves. From section IIC, we know that the average displacement induced by the function ξw is equal to dα(w) with ds < dl, while no constraints emerged concerning the direction of this displacement as a function of word type. Therefore, to average the functions Ξi over all word lists of one type, we replace ξw with the function ¯ξα(w) defined by ¯ξα(x) = x + dαe, where e is a unity vector randomly chosen from a suitable distribution Ω0[e]. We can thus write Ξi(y0) ∼ y0 +(cid:80)j (cid:42) pretrieval(cid:16)(cid:12)(cid:12)(cid:12) pi((cid:126)α; y0, x0) = where Ω(e1, . . . , ei) =(cid:81)i k=1 Ω0[ ek]. i(cid:88) k=1 k=1 dαk ek, and eq. (8) becomes (cid:12)(cid:12)(cid:12)(cid:17)(cid:43) Ω dαk ek + y0 − x0 15 (9) The initial positions y0 and x0 are not measurable quantities. In eq. (9), therefore, they must be averaged over through two suitable distributions χ(y0) and ψ(x0), yielding i(cid:88) (cid:12)(cid:12)(cid:12)(cid:17)(cid:29) (cid:28) pretrieval(cid:16)(cid:12)(cid:12)(cid:12) k=1 dαk ek and p(yx) := pretrieval(cid:0)(cid:12)(cid:12)y − x(cid:12)(cid:12)(cid:1). dαk ek + y0 − x0 k=1 χ,ψ,Ω pi((cid:126)α) = where yi :=(cid:80)i The function we need to average may be rewritten as (cid:28) (cid:29) p(yix0) = (10) χ,ψ,Ω p(yix0) = P(cid:2)i1 = i(cid:3) + P(cid:2)i1 = j(cid:3) p(yiyj) (cid:88) j(cid:54)=i where P(cid:2)i1 = i(cid:3) is the probability that the i-th memory, yi, will be the first one to be found, and(cid:80)N P(cid:2)i1 = i(cid:3) < 1. Substituting eq. (11) into eq. (10), we find (cid:43) χ,ψ p(yiyj) (cid:42)(cid:10)P(cid:2)i1 = i(cid:3)(cid:11) (cid:10)P(cid:2)i1 = j(cid:3)(cid:11) (cid:88) pi((cid:126)α) = χ,ψ + (12) (11) i=1 j(cid:54)=i Ω The distribution ψ refers to the state x0 of the system after the so-called retention interval, during which the subject freely elaborates the information gathered during presentation. A full understanding of such elaboration would require modeling the free motion of this system, but we have only been able to markovianize the equations of motion during a progressive searching task or in the presence of a driving input. Thus, a reasonable ansatz is necessary. Neglecting recency effects, we can suppose ψ to contain N similar peaks at the locations y1, y2, . . . , yN explored during presentation. In this picture, the dependence of(cid:10)P[i1 = i](cid:11) on the index i will be negligeable, so we can approximate value p0. Substituting this into eq. (12), and rewriting the argument of pretrieval, we find (cid:34) (cid:42) pretrieval(cid:16)(cid:12)(cid:12)(cid:12) max(i,j)(cid:88) (cid:88) j(cid:54)=i pi((cid:126)α) = p0 1 + k=min(i,j)+1 Ω (cid:69) (cid:68)P[i1 = i] (cid:12)(cid:12)(cid:12)(cid:17)(cid:43) dαk ek χ,ψ (cid:35) χ,ψ with a constant (13) where Ω is now the product of j − i copies of the distribution Ω0, and the value of the summand depends solely on the number of long and short words located between word i and word j. Using pretrieval(0) = 1 to include the j = i term into the sum, and noticing that the labels 16 of the ek's are interchangeable, we can finally rewrite eq. (13) as (cid:88) pi((cid:126)α) = p0 π(Sij, Lij) where we have introduced the quantities j max(i,j)(cid:88) min(i,j)+1 Lij := π(m, q) := 1(αk = l) (cid:28) pretrieval(cid:16)(cid:12)(cid:12)(cid:12)ds m(cid:88) k=1 Sij := i − j − Lij (cid:12)(cid:12)(cid:12)(cid:17)(cid:29) Ω q(cid:88) k=1 ek + dl hk (14) (15) (16) and the unit vectors gk, hk are independently distributed according to m + q copies of the distribution Ω0. For pure lists, eq. (14) becomes ppure i,α := pi(α, α, . . . , α) = p0 (cid:20) 1 + (cid:16) N−i(cid:88) i−1(cid:88) (cid:17) π(cid:0)δαsh, δαlh(cid:1)(cid:21) + h=1 h=1 (17) C. Coexistence of word length effects As mentioned above, the probability of recalling the i-th word of the list is equal to the probability of retrieving the i-th memory, multiplied by the factor qwi. In the previous section, we averaged the retrieval probability over all words of the same type; similarly, we must now average the verbalizability qx, which we take to be distributed independently of the word structure {vx}x. Defining qα as the average verbalizability of words of type α, we obtain the full recall probability Pi(α1, . . . , αN ) = qαi pi(α1, . . . , αN ) (18) As mentioned in the Introduction, the classical WLE is the experimental fact that pure lists made of shorter words are easier to remember. Of course such behavior may always 17 be prevented, in principle, by making the verbalizability ratio qs/ql sufficiently low. Yet, if the reported inversion of the WLE exists within this model, it must rely on the opposite requirement -- namely, that the verbalizability ratio is sufficiently high. The questions we are therefore supposed to answer are: first, whether there exists for this model a range of values of qs/ql where both the classical and the inverse WLE occur; second, under what conditions on the parameters this may happen, and whether such conditions are relevant to current experiments. Let us describe the typical experimental situation. In the experiments, word lists are generated by drawing words at random from a vocabulary W. scenario, this vocabulary will contain Ws short words and Ws long ones. We may consider, therefore, an ensemble of lists of length N , where each word has a probability γ := Wl of being long, and a probability 1 − γ of being short. In a double-word length Ws+Wl The recall probability for the i-th word of the list can be averaged over all lists whose i-th word is of type α, yielding pi,α(γ) := (cid:104)pi((cid:126)α)(cid:105)γ. Substituting eq. (14), this becomes: (cid:34) (cid:88) (cid:68) S,L≥0 0≤S+L≤N−i (cid:69) (cid:88) (cid:68) S,L≥0 0≤S+L≤i−1 (cid:35) (cid:69) pi,α(γ) = p0 π(S, L) + γ π(S + δαs, L + δαl) (19) γ In the first sum, which is the contribution from j ≥ i in eq. (18), S and L stand for the number of short or long words in positions k such that i + 1 ≤ k ≤ j; in the second sum (the contribution from j < i) S and L stand for the number of short or long words in positions k such that j + 1 ≤ k ≤ i. In eq. (19), the notation(cid:10) . . .(cid:11) γ has come to denote an averaging over S, L performed for each separate value of S + L and summed together. This is done through the binomial distribution (cid:18)L + S L (cid:19) γL(1 − γ)S Φ(S, L) = (20) Given that the total recall probability is Pi,α(γ) = qαpi,α(γ), the standard WLE effect (Pi,s(0) > Pi,l(1)) will occur if ql qs will occur if Pi,s(γ) < Pi,l(γ), that is, if ql qs effects coexist if θinv < ql qs rewritten as < θcl, where θcl = pi,s(0)/pi,l(1), and the inverse effect > θinv, where θinv = pi,s(γ)/pi,l(γ). The two < θcl , which can only happen if θinv < θcl. This condition can be pi,l(1)pi,s(γ) < pi,s(0)pi,l(γ) 18 (21) Now, it can be seen that pi,α(γ) a decreasing function of γ, because by increasing γ one transfers weight from the first to the second argument of π inside both terms of eq. (19), which reduces the value of the function π. Hence, we have pi,l(1) < pi,l(γ) and pi,s(γ) < pi,s(0), from which it follows that the inequality (21) is identically satisfied. We conclude that, in this model, the WLE can undergo an inversion for any γ ∈]0, 1[, as sketched in Figure 1. D. Formulas for recall probabilities: slow-diffusion regime Consider a list of the type (α1, . . . , αN ), where αi ∈ {s, l}, containing L words and S short ones. The trajectory during presentation is illustrated in Figure 8. The system begins from a random position y0, and at each new word wi of type αi, its position is shifted forth by the operator ξwi. The distance travelled at step i is, in the mean field approach, equal to dαi. So a memory produced by the presentation of a short word will be formed in the vicinity of the latest memory (that is, within a distance of order ds) whereas a memory produced by the pre- sentation of a long word will be formed at a longer distance (of order dl) from the memory preceding it. Consequently, memories are divided into clusters separated by a distance of order dl from each other. Each cluster spreads over a width of order ds. FIG. 8: Trajectory during the presentation of a mixed list. The list structure corresponding to the example is lsslsssslsssslss, where l stands for long words and s for short ones. The longer jumps correspond to the presentation of long (localized) words. These memory clusters correspond to different "segments" of the list. Call {li}L i=1 the index values corresponding to long words within a given list. If l1 = 1, the segments are 19 (cid:126)si = (wli, wli+1, . . . , wli+1−1) for i = 1, . . . , L − 1, and (cid:126)sL = (wlL, wlL+1, . . . , wN ). If w1 is short, there is an additional segment (cid:126)s0 = (w1, w2, . . . , wl1−1), and the number of segments is L + 1. Call ai the length of segment (cid:126)si. All the ai's must be positive except a0, which may be null, and(cid:80)L i=0 ai = N . The direction in which the trajectory moves at each step is defined by the unknown distribution Ω0, which depends on the details of the word structure {vx}x. For a generic choice of Ω0, clusters formed at longer time intervals from each other will lie further apart processes: a clustering process and a diffusion process. in semantic space. Hence, the trajectory during presentation is the composition of two Notice that in eq. (13) the argument of pretrieval is of the order of (cid:112)Sijd2 the list is short enough (that is, if pretrieval((cid:112)Sd2 l . If √ Sds) ∼ √ pretrieval(ds) and pretrieval(dl) (cid:28) pretrieval( Sds)), the summand has two orders of magnitude: one of the order of pretrieval(dl) = pl and one of the order of pretrieval(ds) = ps. This s + Ld2 l ) ∼ pretrieval(dl), pretrieval( s + Lijd2 corresponds to the fact that the diffusion process is slow on the scale of the trajectory during presentation. In this regime, formula (13) for pi((cid:126)α) may be estimated by replacing the summand with pl whenever there is at least one long word between min(i, j) + 1 and max(i, j), and with ps otherwise. This is equivalent to approximating the matrix elements π(m, q) of eq. (16) with π(1, 0) if q = 0 and m > 0, and with π(0, 1) if q > 0, thus ignoring the dependence of the retrieval process on the distance between segments of the list. Call yi1 the first memory to be retrieved. The conditional retrieval probability for memory yj is pretrieval(yj − yi1). In the slow-diffusion limit, this is of the order of ps if memories yi1 and yj belong to the same cluster, and is of the order of pl otherwise. Averaging over the first retrieval, we find (cid:104) p(i) = p0 1 + (ci − 1)ps + (N − ci)pl (cid:105) where ci is the length of the segment of the list to which word i belongs. The average recall probabilty P α for words of type α will thus be equal to: (22) (23) (24) L(cid:88) i=0 Ps = qsp0 S (cid:16) (cid:17)(cid:104) (cid:104) L(cid:88) i=1 ai − 1 + δ0i 1 + (ai − 1)ps + (N − ai)pl Pl = qlp0 L 1 + (ai − 1)ps + (N − ai)pl (cid:105) (cid:105) Defining µ = N−a0 L and ∆ = 1 L (cid:80)L i=0 a2 i , we can rewrite this as (cid:104) Ps = qsp0 N pl − ps + 1 + (cid:104) L N − L (∆ − µ)(ps − pl) N pl − ps + 1 + µ(ps − pl) Pl = qlp0 20 (25) (26) (cid:105) (cid:105) All the dependence on the ordering of words in the list, therefore, enters the recall prob- abilities through the parameteres µ and ∆. The values of these parameters are shown in table I for simple lists. TABLE I: Values of the observables µ, ∆, and Ps/Pl for simple lists List Structure l . . . l l . . . l s l . . . l l l l . . . S s . . . s (cid:124)(cid:123)(cid:122)(cid:125) (cid:124)(cid:123)(cid:122)(cid:125) (cid:124)(cid:123)(cid:122)(cid:125) l s . . . s s . . . s M m l s . . . s . . . l s . . . s (cid:124)(cid:123)(cid:122)(cid:125) m (cid:124)(cid:123)(cid:122)(cid:125) l s . . . s m µ 1 1 ∆ 1 1 + S2 N−S Ps/Pl -- (N−S)pl+(S−1)ps+1 (N−1)pl+1 qs ql N N−1 N − M N +2 N−1 N 2 − 2M (N − M ) qs ql N 2ps−2N ps+ps+N−1+M (2M−2N +1)(ps−pl) N pl+ps+1 N 2pl+N−2N pl+ps−1 qs ql (N−1)[(N−M )ps+M pl−ps+1] m + 1 (m + 1)2 qs ql For a pure list, entirely composed of words of type α, eqs. (25) and (26) become: (cid:2)1 + (N − 1)pα (cid:3) P pure α = qαp0 (27) Data for mixed lists may be interpolated with formulas (25) and (26) to test the theory and fix the values of the internal parameters. Finally, let us look at the range of occurrence of the WLEs in the slow-diffusion regime. The classical WLE (P pure s > P pure l ) emerges for ql qs < θcl where θcl = 1 + (N − 1)ps 1 + (N − 1)pl (28) The ratio Pl/Ps for mixed lists can be estimated as follows. Given a fixed value of L < N , µ ranges between µmin = 1 (for a0 = N − L) and µmax = N/L (for a0 = 0). The minimum 21 value of ∆ is obtained by starting the list with a short word and having N − (L + 1)(cid:100) N−L−1 L+1 (cid:101) L+1 (cid:101)) − N segments of 1 + (cid:98) N−L−1 L+1 (cid:99) segments of 1 + (cid:100) N−L−1 words. The maximal ∆ is obtained by setting a0 = 0 and lumping all the short words into one segment: ∆max = (N−L+1)2+L−1 L+1 (cid:101) words and (L + 1)(1 + (cid:100) N−L−1 L . Substituting ∆max and µmin in (25), (26), one obtains a strict upper bound on the ratio Ps/Pl: Ps Pl < qs ql θinv, where θinv = L2(ps − pl) + N L(pl − 2ps) + psN 2 + (N − L)(ps − 2pl + 1) (N − L)[1 + (N − 1)pl] (29) Once again, the inverse WLE (Ps/Pl < 1) will emerge under the sufficient condition > θinv; hence, the possibility that the classical and inverse effects may coexist requires ql qs θcl > θinv. If we rewrite this inequality by means of eq. (28) and (29), the "microscopic" probability parameters ps and pl cancel out and we obtain the general condition L2−LN +2(N −L) < 0, which is identically satisfied for any L > 2. It follows that, in the slow-diffusion regime, the system displays an inversion of the WLE for all mixed lists containing more than two long words. This statement, unlike the conclusions of the previous section, is not only true on average, but holds true regardless of the order in which short and long words are arranged within the list. E. Formulas for recall probabilities: fast-diffusion regime In the previous section we have considered the case where the diffusion process was much slower than the clustering process -- which translates into upper bounds on the order of magnitude of S and L. Notice that we defined these bounds in terms of the parameters ds and dl, whose value may vary from subject to subject. Therefore, the very same list may be experienced in a fairly stationary regime by one subject, and in a regime of fast diffusion by a more easily distracted one. Here, we will consider the opposite case of very fast diffusion, defined as the regime where pretrieval(ds + x) goes down fast on the length scale of ds, so the matrix elements of π decay quickly as the value of the indices grows, and the recall dynamics is dominated by the contiguity effect. If a long word causes a lesser diffusion than two short words, we may neglect all but the top-left elements of the π-matrix: π(0, 0) = 1, π(1, 0) = ps, and 22 π(0, 1) = pl. Otherwise, we can work in the lowest approximation by neglecting also pl. Eq. (19) becomes (cid:104) 1 + ps + pα − (ps − pl)γ (cid:105) Pα(γ) ∼ p0qα (30) where I dropped the pedix i because all dependence on i vanishes as long as i is neither 1 nor N . The thresholds for the verbalizability ratio are θcl = 1 + 2ps 1 + 2pl θinv = 1 + 2ps + (pl − ps)γ 1 + ps + pl + (pl − ps)γ (31) and for θinv < ql qs < θcl, eq. (30) is nothing but a linear version of Figure 1. When experiments yield a near-linear curve Pα(γ), thus, it may be taken as a sign that the system is operating in a very fast diffusion mode. When the observed curve is nonlinear, one must infer that terms of the type π(h, 0) for h > 2 are relevant, and a polynomial interpolation can be performed, by truncating the sums in eq. (19). The crossover toward linearity of the curves Pα(γ) may be explored by tuning experi- mentally the amount of diffusion in the system, as we will see in section V. IV. DATA ANALYSIS A. The Data The foregoing analysis has proven that, within this model of verbal perception, the recall process can display consistently both the classical and the inverse WLE, as observed in the experiments. There may be, however, other mechanisms leading to a similar prediction. Katkov et al, for instance, propose a lexical explanation for both the direct and inverse WLE, based on possible differences in the long-term neural representation of long and short words (Katkov et al., 2014). In order to tease out which mechanism is really responsible for the effects, one needs to extract further information from the data, going beyond the computation of mere recall probabilities. 23 I will do so by using data from PEERS (Penn Electrophysiology of Encoding and Retrieval Study), a large study conducted at the University of Pennsylania and devoted to assembling a database on the electrophysiological correlates of memory (Lohnas and Kahana, 2013). The sample I have considered corresponds to Experiment I of PEERS. It includes data from trials on 156 college students (age range: 18−30) and on 38 older adults (age range: 61−85 years). In each trial, 16 words were presented one at a time on a computer screen. Each word was drawn from a pool of 1638 words with length varying between S = 1 and S = 6 syllables. The word-length distribution was peaked at S = 2, reflecting that of a typical English lexicon (but not of a typical English corpus). Each item was kept on the screen for 3000 ms, followed by an interstimulus interval of 800−1200 ms. After the last item in the list, there was a delay of 1200− 1400 ms, after which the participant was given 75 s to attempt to recall aloud any of the just-presented items. Multiple trials were performed on each subject, summing up to 3744 trials for the students sample and to 912 for older adults. For more details on the experimental procedure, see Healey and Kahana, 2016. Because the word lengths involved are more than two, the formulas we derived in the previous section may not be applied verbatim. Nonetheless, two key consequences of the theory afford a direct comparison with the data. B. Recall by contiguity We say that a word is recalled "by contiguity" when its recall occurs immediately after the recall of a word contiguous to it within the list. In the model we have developed, no matter how high the amount of diffusion, recall by contiguity will be more frequent for short words than for long words. This follows from the fact that two consecutive short words belong necessarily to the same segment, whereas a short word and a long word, though contiguous, may belong to different segments, and two long words are sure to belong to different segments. While this prediction has been derived in a two-length model, it generalizes immediately to models with multiple lengths. We have proven in section II that shorter words are characterized by a shorter reaching distance; hence, memories formed by shorter words have a higher chance of being located in the proximity of memories formed by contiguous words. Contiguity will play therefore a stronger role in the recall of a shorter word: the shorter the words involved, the higher the chances of recall by contiguity. 24 Pcon S FIG. 9: Probability of recall by contiguity for words with S syllables (blue dots); probability of recall by contiguity (black); and mean probability of recall by contiguity (red). All values have been computed from PEERS data. Let us call Pcon(w) the probability that a given word w, if recalled, will be recalled by contiguity. The theory predicts that Pcon should be larger for shorter words. If so, computing Pcon from the data and averaging it over all words of the same length would yield a decreasing curve Pcon(S). In Figure 9, Pcon is shown for all words having the same number of syllables (blue dots). The data have been aggregated from all trials. The distribution of Pcon (black line) is wide for all word lengths; nevertheless, the mean probability of recall by contiguity (shown in red) decreases monotonically with the number of syllables, in agreement with the theory. The 1638-element wordpool used in PEERS contains only four 4 five-syllable words, and a single 6-syllable word ("encyclopedia"). Hence, the statistics for these two lengths may not be considered reliable. While Figure 9 refers to Experiment 1 of PEERS, the robustness of the effect has been 12345600.10.20.30.40.50.60.7PconPconhistogrammean(Pcon) 25 checked for by repeating the analysis on the databases from Experiments 2 and 3 (see Lohnas et al., 2015). The decreasing trend has proven invariant across databases. C. Distribution of Jumps Let (i1, i2, . . . , iM ) be the serial positions of the words recalled by the subject during a certain trial. At each step n in the recall process, the system performs a serial-position jump of a magnitude δn = in − in+1. In our terminology, longer jumps will require the exploration of a larger portion of semantic space. Therefore, the distribution of jumps should be monotonously decreasing as a function of their size (as appears indeed from the PEERS data, Figure 10). Pt δ FIG. 10: In black, distribution of serial-position jumps in the recall process as a function of jump size δ, computed from PEERS data. In colors, distribution of jumps for a fixed size b of the word-length barrier, shown in the legend. As the lists contain 16 words, δ ranges between 1 and 15. From our analysis of the two-length scenario, we know that these jumps cover greater distances in semantic space if the memory created by the word of departure and the memory created by the word of arrival belong to different clusters. Since it is long words that have 246810121400.050.10.150.20.250.30.350.40.450.5b=1b=2b=3b=4 the ability to break clusters, a long word located in a position k such that i < k ≤ j plays effectively the role of a recall barrier, hindering direct transitions between wi and wj. 26 FIG. 11: To the left, depiction of the hyerarchical segmentation of a 16-word list. Word lenght is encoded in colors as shown in the vertical sidebar. At each level of segmentation, clusters are displayed as horizontal bars covering the given portions of the list. To the right, a plot of the mean cluster length (blue) and of the number of clusters per list (red), as a function of the hierarchical level S of clusters, computed from PEERS data by averaging over all trials. S Similarly, in experiments with multiple word lengths, longer words create memories at longer distances. As clustering in semantic space will occur then over multiple length scales, clusters assume a hyerarchical structure. So does the segmentation of the list, with mod- erately long words marking off smaller segments within the larger segments delimited by longer words. This is shown in Figure 11, together with the average length of clusters as computed from the data, and the average number of clusters per list. The probability for the transition wi → wj (a jump of size i− j) will be affected mainly by the hyerarchical level of the clustering involved. Since longer words break clusters at a higher level, we conclude that the effect of clustering on such a jump will be controlled by the length of the longest word located between the positions i and j: max{L(wj+1), L(wj+2), . . . , L(wi)} if i > j max{L(wi+1), L(wi+2), . . . , L(wj)} if j > i bij = (32) where L(w) is the length of word w and, once again, the lowest index has been excluded for the same reasons why it was not counted in eq. (13) of section III. 1234560246810121416cluster lengthnumber of clusters per list123456 27 FIG. 12: Serial-position jump from word 8 to word 3 during the recall of a nine-word list (with words from the PEERS wordpool). The red bars describe word length: the "barrier word" for this transition is marked in yellow. We will call bij the "barrier size" for the i → j transition. The role played by this quantity is depicted in Figure 12. If the clustering mechanism does exist, it will enhance the transition probability for small values of bij, and suppress it for higher values. To test this prediction, we must first study the problem in the absence of clustering, calculating the transition probability Pt(δ) for jumps of size δ. The resulting formula must be evaluated using the experimental values of parameters, and can then be compared with the values of Pt(δ) extracted from the data. If the clustering occurs, we will observe an enhancement of Pt for small values of the barrier size b and a suppression at high values. Let us suppose that the clustering does not occur. The system then will not "feel" the word-length barriers, and the probability Pt of a jump will be independent on the size of the barrier to be overcome. Conversely, the probability distribution of barriers over the jumps performed by the system will depend only on the jumps' size, and on the probability distribution of word length with the list. The probability that a jump of size δ involves a word-length barrier b is equal to the probability that the longest word out of a random sequence of δ words has length b. This is equal to ZebraSidingTubeRefrigeratorNewspaper (cid:33)δ f (L) (cid:32) b(cid:88) L=1 (cid:32) b−1(cid:88) L=1 − (cid:33)δ f (L) Pt(bδ) = 28 (33) where f (L) is the word-length distribution with the lists. Formula (33) describes the no-clustering barrier-size distribution, that is, the distribution of barrier sizes for jumps of a given length if the jumps are not influenced by the barriers. I have evaluated formula (33) using values of f (L) computed directly from the lists recorded in the PEERS database. In Figure 13 (see the end-page), the results are compared with the values of Pt(bδ) extracted directly from data. The no-clustering curves are dashed, the experimental curves solid. The effect of clustering is apparent. All experimental curves reveal an enhancement of the transition probability for small values of b, and a suppression at large values. Morevoer, the effect persists across the whole range of possible jump sizes. We would expect that the influence of clustering should be small for jumps shorter than the typical size of the clusters, and grow as the jumps get longer. This is indeed what we notice in Figure 13. The influence of clustering can be observed already for jumps of size one, with an enhancement of probability for small barriers and suppression for larger barriers. Yet, the clustering becomes more prominent as the jumps get longer. A noticeable increase in the importance of clustering is observed after δ = 6. This may be due to the fact that the S = 3 clusters are becoming important, as the jumps have gotten longer than the average length of the clusters of the third level (see Figure 11). Since the PEERS lists have length 16, there is no available statistics for values of δ larger than 15. Nonetheless, we can predict that this growth will saturate when the probability of meeting a maximal bareer approaches near-certainty. For jumps longer than that, b is no longer a controlling parameter; from that point on, the continued suppression of recall will derive from the need to find the next memory multiple clusters away. V. FURTHER TESTS A. Predictions on average recall probabililties Most free-recall experiments performed so far have employed either a wordpool with an arbitrary number of syllables, as in PEERS, or a pool characterized by a fixed length that 29 is usually 2 syllables (as in the Toronto Wordpool, see Friendly et al., 1982). Further tests on this theory, however, would be easiest to perform through experiments in which the words presented for recall are drawn from either of two sets: a set of very long words (e.g. number of syllables S = 4) and a set of very short ones (S = 1). When the lists are sufficiently short, the system operates in a slow-diffusion (or "cluster- ing") regime. In this case, three parameters should be kept track of in the experiment: ∆, µ, and the number L of long words in each list (see section III). While the recall probability of individual words depends on the details of the list, the average recall probability for short or long words should be affected by the list structure only through those three parameters, which completely characterize the list for the purposes of this type of experiment. It is then feasible to test qualitatively the following predictions: 1. P s correlates positively with ∆ and negatively with µ. 2. P l correlates positively with µ. These are straightforward consequences of (25), (26). In experiments with words of two lengths, data may be interpolated with formulas (25) and (26) to test the theory and to fix the values of the internal parameters. B. Predictions depending on serial position Serial position effects related to word-length can be isolated experimentally by using word lists constructed on two or more basic templates of segment-structure (cid:126)α = (α1, . . . , αN ), with αi = s, l. The experimenter would present equivalent lists to a number of participants and would correlate the recall probability of words with their positions within each template structure. Large data sets from such experiments should be able to corroborate or to rule out the mechanisms I have described. Competing effects of familiar types, such as primacy and recency, may be easily subtracted. Two simple predictions can be made, easier to check in a slow-diffusion regime but equally valid with faster diffusion: 1. Words from longer segments have a higher recall probability than words of the same type from shorter segments; 2. Successful recall of words from one segment hinders the recall of words from others. 30 The first prediction follows from eq. (22), and should be easy to test. The second prediction must be compared with data by computing the correlation func- tions Cij, that is, the joint probability for the recall of the i-th and j-th words in the list. From there, it is straightforward to obtain the in-segment and cross-segment correlation functions: (cid:42) (cid:43) (cid:42) (cid:43) C in α1α2(d) = Cij i−j=d i,j∈ same segment α(wi)=α1,α(wj )=α2 C cross α1α2 (d) = Cij i−j=d i,j∈ different segments α(wi)=α1,α(wj )=α2 (34) and to verify whether C in α1α2(d) > C cross α1α2 (d), as the theory suggests. We may add to these predictions a third one -- namely, the fact that the long word of each segment is the easiest one to recall. This results directly from the bounds we derived in the previous section on the ratio qs/ql. C. Inter-response intervals A further prediction of the theory regards inter-response intervals -- that, is the time elaps- ing between one recalled item and the next -- whose measurement in free-recall experiments dates back to (Murdock and Okada, 1970). During retrieval from long-term memory, it was shown (Gruenewald and Lockhead, 1980) that clusters occur due to stable semantic associations between objects: a subject who is asked to list some animals, for instance, may recall first a set of farm animals, then a number of house pets, then several birds. The inter-response intervals are shorter within clusters than between clusters. The retrieval of examples in the experiment of Gruenewald and Lockhead depends entirely on the long-term representation of items. In the situation we have described, on the contrary, short-term memory is at play, and the retrieval process has to locate the vanishing traces of a recent experience. Nonetheless, it is easy to see that the time interval elapsing between the retrieval of two memories will be longer between two memories belonging to different clusters, and shorter for memories belonging to the same cluster. Hence, the same is true among items of the list that are successfully verbalized. The inter-response intervals will be longer for the consecutive recall of two words belonging to 31 different segments of the list, and shorter for the consecutive retrieval of two words belonging to the same segment. D. Experiments with varying presentation rate As mentioned in the Introduction, the WLE has been reported in experiments where the time interval between the presentation of consecutive items was a controlled parameter. Experiments on the WLE with rapid presentation of the stimuli were first performed by Coltheart and Langdon (1998), who found the WLE by presenting an item every 114 ms, every 157 ms, and every 243 ms. In (Campoy, 2008), somewhat lower presentation rates were used (between 300 and 400 ms), and again the persistence of the effect was proven over different rehearsal times. Here, I will argue that such experiments may offer an ideal tool to study the crossover between the "diffusive" and the "clustering" regime. Indeed, if we modify the foregoing computation to allow the system to random-walk on its own for a time τ between the presentation of the i-th and i + 1-th items, this will be equivalent to increasing the average distance d travelled between the memory yi generated by word wi and the memory yi+1 generated by word yi+1. This amounts to rescaling time while replacing dl and ds with larger effective distances. The matrix elements of π will decay faster as their indices grow. Therefore, the system will move closer to the diffusive regime. On the other hand, the theory predicts (section IIIE) that the curve Pα(γ) will become linear in the fast-diffusion regime. Hence, by reducing the presentation rate, one should see the two curves in Fig. 1 becoming progressively linearized, at least up to values of τ so large that not only the clustering, but also the contiguity effect breaks down. In this limit, moreover, eq. (30) predicts that the curves Ps(γ) and Pl(γ) will become parallel. Other ways of controlling the crossover between clustering and diffusing regime (e.g. by pharmacological means, or through distractor tasks such as those of Bjork and Whitten, 1974) can be similarly applied. 32 E. Subject-dependent variability of the classical WLE We have just examined some ways of testing the theory that are going to require ad-hoc experiments. Let us conclude by mentioning one type of test that may be performed on already available databases -- namely, data from the experiments performed so far on the classical WLE. It is known that different people employ different strategies in order to exploit the struc- ture of semantic space during memorization and recall (Healey et al., 2014). When it comes to the clustering described above, we expect its strength to be just as dependent on the subject considered. On the other hand, the strength of the classical WLE is also a function of the particular subject. The difference between the recall probabilities of all-short lists and all-long lists, while being positive on average, will vary in magnitude from subject to subject. If the mechanism at the root of the of the WLE is indeed a clustering phenomenon, we expect that it will be stronger for subjects for whom the clustering is stronger. Let (i1, i2, . . . , iM ) be again the serial positions of the words that have been recalled by (cid:80)M−1 the subject during a certain trial. The average serial-position jump between consecutively n=1 in − in+1. For each subject, we can define δsubject as the average of δtrial over all trials performed on him or her. The parameter δsubject recalled words for this trial is δtrial = 1 M−1 may serve as a simple measure of the subject's inclination toward clustering. Indeed, δsubject is close to unity for subjects strongly inclined toward clustering, and δsubject (cid:29) 1 if the subject's tendency toward clustering is low. One way of gaining insight into the likelihood of the theory consists in correlating δsubject with the strength of the classical WLE effects, controlled by ps− pL, where ps and pL are the recall probabilities calculated for pure lists of short and long words. A negative correlation between these two variables would be a strong confirmation that the WLE is indeed due to a clustering phenomenon. VI. INTERPRETATION AND CONCLUSIONS I have proposed a theory of verbal perception, extracted some of its properties, and validated it through a comparison with free-recall data. 33 The theory is based on the notion that a word does not have, in general, a single meaning. A human subject exposed to a stream of verbal input will decide on the meaning of each new word on the basis of both the structure of its vocabulary and the meaning he/she has given to the words preceding it. This also applies to a list of random words, because our mind strives to interpret them as parts of a meaningful discourse. It may be instructive to think of such discourses as "narratives". Common experience tells us that a two-word list is already capable of creating a strong narrative sense (e.g.: picnic, lightning). When a word in the list has no semantic connection to the context created by the words preceding it, the mind perceives a "change of scenery" and assumes that a new narrative is beginning. A list of words is thus perceived as a collection of distinct "stories". When prompted to recall the list, the subject remembers each story as a separate experience, and needs to re-experiences a given story before retrieving the words responsible for creating it. Words that have specific meanings have obviously less probability of fitting into a ran- domly generated story. Otherwise said, the words most likely to break the narrative are those with the highest level of localization in semantic space. We have argued that this correlates positively with word length. Hence, a list of N long words is likely to break into as many one-word stories, whereas a list of short words is more likely to be perceived as a single continuous narrative. Since a single narrative is easier to recall than many unrelated ones, the standard word length effect ensues. The clustering property of short words is at play in mixed lists as well. But its effect is hindered by the presence of long words breaking the narrative. As our analysis has shown, this can lead to an inversion of the WLE, encountered in experimental observations. In this scenario, the behavior depicted by Figure 1 becomes quite logical. By replacing a short word with a long word, one splits the list into a larger number of narratives, which makes every single word in the list (whether short or long) harder to reach during the retrieval process. The interplay between the trajectory of the system during the presentation of lists and the trajectory during the memory test produces a nontrivial spectrum of behaviors, highly dependent both on the structure of lists and on the amount of "diffusion" that interferes with clustering. 34 A telltale symptom of these mechanisms is that a short word is more likely than a longer one to be recalled right after a word contiguous to it within the list. An analysis of data from the PEERS experiments (Healey and Kahana, 2016) confirms this prediction. Another prediction of the theory is that the serial-position jumps performed during recall will be enhanced by clustering for shorter words, and suppressed for longer ones. This is also confirmed by an analysis of PEERS data, across the whole available spectrum of jump lengths. Several directions stand open for experimental and theoretical work on this model. Ex- perimentally, further tests may be obtained by performing the five types of measurements listed in section V. Open directions for theoretical work include: 1. generalizing the recall probability formulas to the case where the word lengths available are more than two; 2. including possible competition between words for the verbalization of a given state; 3. sin- gling out extra effects from the fluctuations around the mean field behavior; 4. accounting for primacy and recency effects; 5. applying the same technique to predicting the genesis of false memories. Finally, by positing a suitable mechanism for spontaneous language production, it would be useful to derive equations linking the underlying word structure to emerging verbal pat- terns, thus providing a direct link between the hidden variables and the observables of the model. I am grateful to Michael J. Kahana and his University of Pennsylvania research group for making their raw data publicly available. I would like to thank heartily people at the Neurotheory Center of Columbia University (Ken Miller, Misha Tsodyks) for hosting me there in the fall of 2015 and for responding with enthusiasm to this theory. Thanks also to Yashar Ahmadian of the University of Oregon, for reading an early draft of the paper and suggesting several improvements. VII. BIBLIOGRAPHY Aurenhammer F., Klein R., and Lee D.T., 2013. Voronoi Diagrams and Delaunay Trian- gulations. Singapore: World Scientific Publishing Company. Baddeley, A.D., Hitch, G., 1974. Working memory. In G.H. Bower (Ed.), The psychology of learning and motivation: Advances in research and theory, Vol. 8, pp. 47-89. New York: 35 Academic Press. Baddeley A.D., Thomson N., Buchanan M., 1975. Word length and the structure of short-term memory. Journal of Verbal Learning and Verbal Behavior, 14:575-589. Baddeley, A.D., 2007. Working memory, thought and action. Oxford: Oxford University Press. Bhatarad P., Ward G., Smith J. Hayes L., 2009. Examining the relationship between free recall and immediate serial recall: similar patterns of rehearsal and similar effects of word length, presentation rate and articulatory suppression. Memory and Cognition 37: 689-713. Binet A. and Henry. V., 1894. La memoire des mots. L'annee psychologique, Bd. I 1:1-23. Bjork, R. A. and Whitten, W. B., 1974. Recency-sensitive retrieval processes in long-term free recall. Cognitive Psychology, 6: 173189. Campoy G., 2008. The effect of word length in short-term memory: Is rehearsal neces- sary? Quarterly Journal of Experimental Psychology, 61:5, 724-734. Campoy, G., 2011. Retroactive interference in short-term memory and the word-length effect. Acta Psychol. 138, 135-142. Coltheart, V., Langdon, R., 1998. Recall of short word lists presented visually at fast rates: Effects of phonological similarity and word length. Memory and Cognition, 26, 330342. Elts, J., 1995. Word length and its semantic complexity, in Family and textbooks: 115-126. Tartu: University of Tartu. Friendly M., Franklin P.E., Hoffman D., Rubin D.C., 1982. The Toronto Word Pool: Norms for imagery, concreteness, orthographic variables, and grammatical usage for 1,080 words. Behavior Research Methods And Instrumentation, Vol. 14(4), 375-399. Fucks W., 1956. Die Mathematischen Gesetze der Bildung von Sprachelementen aus Ihren Bestandteilen, Nachrichtentechnische Fachberichte 3:7-21. Greenberg, J., 1966. Universals of language. Cambridge, MA: MIT Press. Gruenewald, P.J. and Lockhead, G.R., 1980. The free recall of category examples. J. Exp. Psychol. [Hum- Learn]. 6, 225-240. Grzybek P., 2007. History and methodology of word-length studies, in "Contributions to the Science of Text and Language", Dordrecht: Springer, pp. 15-90. Haspelmath, M., 2006. Against markedness (and what to replace it with). Journal of 36 Linguistics, 42 (01), 25-70. Healey M., Crutchley P., Kahana MJ., 2014. Individual differences in memory search and their relation to intelligence. J Exp Psychol 143: 15531569. Healey, M. K. and Kahana, M. J., 2016. A four-component model of age-related memory change. Psychological Review, 123(1), 23-69. Howard, M. W. and Kahana, M. J., 2002a. A distributed representation of temporal context. Journal of Mathematical Psychology, 46, 269-299. Howard, M. W. and Kahana, M. J., 2002b. When does semantic similarity help episodic retrieval? Journal of Memory and Language, 46, 85-98. Hulme, C., Suprenant, A. M., Bireta, T. J., Stuart, G., and Neath, I., 2004. Abolishing the word-length effect. J. Exp. Psychol. Learn. Mem. Cogn. 30, 98-106. Jalbert, A., Neath, I., Bireta, T. J., and Surprenant, A. M., 2011. When does length cause the word length effect? J. Exp. Psychol. Learn. Mem. Cogn. 37, 338-353. Kahana M. J., 1996. Associative retrieval processes in free recall. Memory and Cognition 24:103-9. Kahana M. J., 2012. Foundations of Human Memory, Oxford University Press. Katkov M., Romani S., Tsodyks M., 2014. Word length effect in free recall of randomly assembled word lists. Frontiers of Computational Neuroscience 8:129. Klare, G.R., 1988. The formative years. In B.L. Zakaluk and S. J. Samules (Eds.), Readability, its past, present and future, Newark, Delaware: IRA, pp. 14-34. Kruglanski A.M. and Tory Higgin E., 2007. Social Psychology: Handbook of Basic Principles, The Guilford Press; Second Edition edition (p.642). Lohnas, L.J., Polyn S.M.,Kahana, M.J. (2015) Expanding the scope of memory search: Modeling intralist and intralist effects in free recall, Psychological Revier, Vol. 122, No. 2, 337363. Lohnas, L. J. and Kahana, M. J. (2013). Parametric effects of word frequency effect in memory for mixed frequency lists. Journal of Experimental Psychology: Learning, Memory, and Cognition, 39, 19431946. Mahowald K., Fedorenko E., Piantadosi S.T., Gibson E., 2012. Info/information theory: speakers actively choose shorter words in predictable contexts, Cognition, 126: 313-318. Lewis, M. L. and Frank M.C., 2016. The length of words reflects their conceptual com- plexity. Cognition 153: 182-195. 37 Mikk. J., Heli U., and Elts J., 2001. Word length as an indicator of semantic complexity, in Text as a linguistic paradigm: levels, constituents, constructs, Festschrift in honour of Ludek Hrebcek. Trier, 187-195. Miller J.F., Weidemann C.T., Kahana M.J., 2012. Recall termination in free recall. Memory and Cognition 40: 4, Pages: 540 - 550. Murdock B-B, 1960. The immediate retention of unrelated words. Journal of Experi- mental Psychology 60:222-234. Murdock B. B., 1962. The serial position effect of free recall. Journal of Experimental Psychology, 64(5)M:482-488. Murdock B.B. and Okada R., 1970. Interresponse times in single-trial free recall. Journal of Experimental Psychology 86:263-267. Neath I., Bireta T.J., Surprenant A.M., 2003. The time-based word length effect and stimulus set specificity, Psychonomic Bulletin Rev. Jun;10(2):430-4. Neath I., Brown G. D. A., 2006. SIMPLE: further applications of a local distinctiveness model of memory, in The Psychology of Learning and Motivation, ed. Ross B. H., editor. (San Diego, CA: Academic Press), 201-243. Piantadosi S. T., Tily H. and Gibson E., 2011. Word lengths are optimized for efficient communication, Proceedings of the National Academy of Sciences, 108, 9:3526. Piantadosi S. T., Tily H. and Gibson E., 2011B. Reply to Reilly and Kean: Clarifications on word length and information content, Proceedings of the National Academy of Sciences, 108, 20: E109. Polyn, S. M., Norman, K. A., and Kahana, M. J., 2009a. A context maintenance and retrieval model of organizational processes in free recall. Psychological Review, 116, 129-156. Polyn, S. M., Norman, K. A., and Kahana, M. J., 2009b. Task context and organization in free recall. Neuropsychologia, 47, 2158-2163. Roberts W.A., 1972. Free recall of word lists varying in length and rate of presentation: a test of total-time hypotheses. Journal of Experimental Psychology 92:365-372. Romani S., Pinkoviezky I., Rubin A., Tsodyks M., 2013. Scaling laws of associative memory retrieval. Neural Computation 25:2523-2544. Russo R. and Grammatopoulou N., 2003. Word length and articulatory suppression affect short-term and long-term recall tasks. Memory and Cognition 31:728-737. Sederberg, P. B., Howard, M. W., and Kahana, M. J., 2008. A context-based theory of 38 recency and contiguity in free recall. Psychological Review, 115(4), 893-912. Standing L., 1973. Learning 10.000 pictures. Quarterly Journal of Experimental Psy- chology 25:207-222. Tehan G. and Tolan G.A., 2007. Word length effects in long-term memory. Journal of Memory and Language 56:35-48. Xu Zhan and Li Bi-Qin, 2009. The Mechanism of Reverse Word Length Effect of Chinese in Working Memory. Acta Psychologica Sinica, Vol. 41 Issue (09): 802-811. 39 δ = 2 δ = 4 δ = 6 δ = 8 δ = 10 δ = 12 δ = 14 δ = 1 δ = 3 δ = 5 δ = 7 δ = 9 δ = 11 δ = 13 δ = 15 FIG. 13: Probability Pt of serial-position jumps in the recall process, plotted as a function of the size b of word-length barriers, for all possible values of jump size δ. The red curves have been calculated in the absence of clustering, as per formula (33). The blue curves are extracted from PEERS data. Clustering manifests as an enhancement of transition probability for low barriers (small b) and as a suppression where larger barriers are encountered. 12345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.512345600.5
1507.07422
1
1507
2015-07-24T01:05:16
Analysis of Pain Hemodynamic Response Using Near-Infrared Spectroscopy (NIRS)
[ "q-bio.NC" ]
Despite recent advances in brain research, understanding the various signals for pain and pain intensities in the brain cortex is still a complex task due to temporal and spatial variations of brain hemodynamics. In this paper we have investigated pain based on cerebral hemodynamics via near-infrared spectroscopy (NIRS). This study presents a pain stimulation experiment that uses three acupuncture manipulation techniques to safely induce pain in healthy subjects. Acupuncture pain response was presented and hemodynamic pain signal analysis showed the presence of dominant channels and their relationship among surrounding channels, which contribute the further pain research area.
q-bio.NC
q-bio
The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 ANALYSIS OF PAIN HEMODYNAMIC RESPONSE USING NEAR-INFRARED SPECTROSCOPY (NIRS) Raul Fernandez Rojas1, Xu Huang1, Keng Liang Ou2, Dat Tran1 and Sheikh Md. Rabiul Islam1 1Faculty of Education, Science, Technology and Mathematics, University of Canberra, Australia 2College of Oral Medicine, Taipei Medical University, Taiwan ABSTRACT Despite recent advances in brain research, understanding the various signals for pain and pain intensities in the brain cortex is still a complex task due to temporal and spatial variations of brain haemodynamics. In this paper we have investigated pain based on cerebral hemodynamics via near-infrared spectroscopy (NIRS). This study presents a pain stimulation experiment that uses three acupuncture manipulation techniques to safely induce pain in healthy subjects. Acupuncture pain response was presented and Haemodynamic pain signal analysis showed the presence of dominant channels and their relationship among surrounding channels, which contribute the further pain research area. KEYWORDS Near Infrared Spectroscopy, Non-invasive Brain Imaging, Hemodynamic, Acupuncture 1. INTRODUCTION Pain is a subjective experience and every individual experiences it in different ways. Pain perception can be range from emotional pain such as the loss of a love one to physical pain such as a broken leg. In addition, pain sensation as the product of dedicated neural mechanisms continues to be a topic of debate. But what is pain? It seems to be very hard to define. According to Perl [1] “With the benefit of the past two centuries of scientific work and thought, can one define pain? Considering the evidence, it seems reasonable to propose pain to be both a specific sensation and an emotion, initiated by activity in particular peripheral and central neurons. Pain shares features with other sensations, but the strong association with disposition is special.” There is a number of publications about pain research [2-4] to answer questions such as: how to interpret it, how to control it, or how to use it to indicate human body health status. In this paper we focus on the instance that an unconscious patient under medical examination cannot express level of pain experience in that moment. For this reason, this research tries to find a relationship between the pain intensity and the hemodynamic response to understand the patient’s feelings. By finding this “silent communication” between pain and brain hemodynamics, medical services will be able to provide a better solution for such pain sensation. Near infrared spectroscopy (NIRS) is employed in our experiments. NIRS can offer simultaneous measurements from dynamic changes of Oxyhemoglobin (HbO) and Deoxyhemoglobin (HbR) in the brain cortex with a reasonable temporal resolution (less than one second) and spatial resolution (less than 3 cm), while Total haemoglobin (HbT) is obtained by the difference between HbO and HbR. Our experimental study makes use of changes in HbO to represent neural activity DOI : 10.5121/ijma.2015.7203 31 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 in the human brain. In addition, acupuncture is also used to induce pain to patients in a safe manner. This paper is organised as follows. Section 2 gives a review of the equipment and experimental setup. The related analysis methods will be discussed responding to the experimental setup. In section 3, the experimental results will be investigated, including dynamic NIRS response to acupuncture stimulation, pain responses of HbO2 and HbR to acupuncture stimulation, pain responses to different acupuncture manipulations, dominant pattern analysis, and direction of pain information. In section 4, conclusions are presented. 2. MATERIALS AND METHODS 2.1. Equipment Hemodynamic measurement was performed using an optical topography system Hitachi ETG- 4000 (Hitachi Medical Corporation). The topography system uses near-infrared spectroscopy (NIRS) to investigate cerebral hemodynamics. Optical topography makes use the different absorption spectra of oxygenated and deoxygenated haemoglobin in the near infrared region. The NIRS system produces 695 and 830 nm NIR signals through frequency-modulated laser diodes. These NIR signals are transmitted to the brain using optical fibre emitters (Figure 1, red circles). Near-infrared light penetrates head tissue and bone, generally it reaches 2 to 3 cm into the cerebral cortex. Once NIR light reaches the cerebral cortex, it is absorbed by haemoglobin, while the non-absorbed NIR signals are reflected to the source, where it is sampled by a high-sensitivity photodiode (Figure 1, blue circles). NIR light between emitters and detectors is sampled at a given time point named channels (Figure 1, numbered squares). Since Oxyhemoglobin (HbO) and Deoxyhemoglobin (HbR) absorb NIR light differently, two wavelengths of light (695 and 830 nm) are used. In this way, it is possible to read these two types of haemoglobin simultaneously. The sample frequency used in this experiment was of 10Hz (10 samples per second). The configuration for this experiment was using two probes of 12 channels to measure neurologic activity of the head (parietal/temporal region). Figure 1 shows the 24 channel configuration used in the study, channels 1 to 12 represent the right hemisphere, while channels 13 to 24 represent left hemisphere. Signals in between channels were computed from the surrounding channels. 2 1 5 4 3 7 6 10 9 8 17 22 12 14 19 24 16 21 11 13 18 23 15 20 Emitter Detector Measurement Channel Figure 1. Channel configuration, right hemisphere (channels 1-12) and left hemisphere (channels 13-24). 2.2. Experiment Setup and Subjects The experiment was designed by the School of Oral Medicine of Taipei Medical University (TMU) in collaboration with the University Of Canberra (UC). In the present study, six healthy individuals (2 females, 4 males) participated in the experiment, aged 25 to 35 years old. All participants provided written consent and the experiment was approved by the Ethics Committee of TMU. The experiment was carried out in the Brain Research Laboratory at TMU in a quiet, temperature (22-24oC) and humidity (40-50%) controlled room. The experiments were done in the morning (10:00am-12:00pm) and each experiment lasted around 30 minutes. Quantitative 32 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 data was collected using the ETG-4000 with the patients sat down in an ergonomic chair near the topography system. The experiment was designed to recognise pain stimulation and pain release in patients through hemodynamic responses. In order to identify the pain through NIRS, acupuncture was used to induce pain stimulation in a safe manner. Brand new acupuncture needles were used for each experiment, and using traditional Chinese acupuncture techniques that were performed by an acupuncturist of TMU Hospital. The puncture point used for stimulation was the “Hegu Point”, located on top of the hand, between the thumb and forefinger. This acupuncture point (acupoint) is known by its property to relieve pain, especially headaches and toothaches. This acupoint is also used to reduce fever, eliminate congestion in the nose, stop spasms, and decrease toothache. A western name for this acupoint is “the dentist’s point” because it can stop tooth pain while moist the throat and tongue. This point was used because it is an area of easy access and the hand can be set aside while the patient is relaxed on the chair. Each patient was punctured on both hands (Figure 2), one hand on a day and the opposite hand on another day; each hand was treated as a separated experiment. The data base was organised of 12 data sets of changes in Haemoglobin files, two sets (right and left hands) per each subject. Figure 2. Patient with acupuncture needle in Hegu point. Three types of pain stimulations (acupuncture techniques) were applied in the experiment, needle insertion (Task 1, T1), needle twirl (Task 2, T2), and needle removal (Task 3, T3). The first type of stimulation (T1) was carried out 30 seconds after the start of experiment and it lasted for 6 seconds. The second type of stimulation (T2) was applied for 30 seconds (rest time) after T1 and it lasted for 10 seconds, this stimulation was repeated three times after 30 second resting time. The last stimulation (T3), was carried out after the third application of task 2, and it lasted for 5 seconds with a 30 seconds recovery time and 15 seconds post-time to finish the experiment. The patients were explained about the acupuncture procedure and experiment. In addition, patients were given a brief explanation of side effects of acupuncture in case they had any symptoms during the experiment; no side effects were reported during and after experiment. After the briefing, the patient was told to sit down on the chair, told to relax and close the eyes to reduce visual evoked stimulation. The optical topography system was set to record after all conditions were met and patient was relaxed. 33 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 2.3. Stages of Experiment Figure 3 exhibits the three types of acupuncture stimulations applied in the experiment. The image shows the changes of HbO, HbR, and (HbT) concentration during the different stages of the experiment; for illustration purposes data was taken from subject 4 in channel 19. The first stage of the study is the pre-time which is used to adapt the patient to the experiment and obtain baseline for the haemoglobin signals, the pre-time lasted for 30 seconds. The second stage (T1, task 1) is started immediately after pre-time stage and it begins the first pain stimulus (needle insertion), and it is done within 6 seconds. After the pain stimulation, patient is given a resting time of 30 seconds (Rt1), which allows the haemoglobin concentration to return to baseline for the next stimulation. The next task (T2) is needle twirl, it lasted 10 seconds and is followed by a 30 second resting time (Rt2); this task was repeated two more times including resting time (Rt3, Rt4) of 30 seconds each. Task three (T3) is started immediately after Rt4 and it lasted for 5 seconds; in this task the needle is removed from the hand. The last recovery time (Rt5) allows the patient to recover from the previous pain stimulus and also haemoglobin signals return to baseline, it lasts for 30 seconds. The last stage in the experiment is the Post-time and lasts for 15 seconds, after the end of this time the equipment stops recording data. ) m m M m * i l ( n b o g o m e H 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 0 Start Finish Rt1 Rt3 Rt4 Pre-time Rt2 Oxy DeOxy Total Rt5 T2 Post-time T1 T2 T2 T3 500 1000 1500 2000 Samples (10/sec) Figure 3. Description of tasks of the experiment, Rt=rest time, T=task. 2.4. NIRS colour representation Figure 4 shows a colour representation of a typical response from a sample of HbO intensity with the 24 channels of subject number 4. The colour representation helped to distinguish pattern trends and focus on areas where the dynamic response was more active. Left figure represents the channel intensities of the right hemisphere, while the right figure represents the channel intensities of the left hemisphere. The colour intensity runs from -0.3 (blue/cold) to +0.5 (red/hot); being the blue colour the lower measurements of HbO, while the red colour represents the higher values of HbO. For visual analysis, values above +0.5 and below -0.3 were represented within this range. 2.5. Data Analysis Signal Analysis was performed by two signal processing techniques. The first method was the cross correlation analysis of raw data to identify time-related channel similarities. While the second method was the Horn and Schunk optical flow (OF) algorithm, this method was used to identify patterns within both hemispheres. Both methods are described as follows. 34 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 05 10 17 22 02 07 12 14 19 24 04 09 16 21 01 06 11 13 18 23 0.3 0.2 0.1 0 -0.1 -0.2 03 08 15 20 Figure 4. Channel intensity, right (channels 1-12) and left (channels 13-24) hemispheres. 2.5.1. Cross Correlation Analysis Data analysis was first performed by comparing channels on both sides of the head. Cross correlation function was computed between channels 1-12 in right probe and 13-24 in left probe. The cross correlation between two input signals (channels) is a kind of template matching. This measure of temporal similarity of two signals was done by computing a time-shifting along one of the input signals. Equation 1 defines the cross correlation between two waveforms x(t) and y(t) as follows: ݎ௫௬ሺ߬ሻ = ࢣ ݔሺݐሻݕሺݐ −߬ሻ ஶିஶ (1) Where ߬ is the time-lag between x(t) and y(t), the value of ݎ௫௬ denotes the difference (lag/lead) between channel signal y(t) and channel signal x(t). The cross correlation value between two channels in the same probe is done from -40 sec to +40 sec at a rate of 10 samples per second. 2.5.2. Optic Flow Analysis The optical flow is defined as the “flow” of pixel values at the image plane in time varying images [5]. Ideally the optical flow corresponds to the motion field, however in some situations the optical flow is not always equal to the motion field. To overcome any problems in the calculation of the optical flow, assumptions have to be defined. Firstly, it is assumed that the surface of the image being observed is flat to avoid variations in brightness due to shading effects. Secondly, the incident illumination is uniform across the surface. Thirdly, it is also assumed that the optical flow varies smoothly and has no spatial discontinuities. We calculate optical flow from the apparent motion at the image plane based on visual perception with the dimensions of the image velocity of such as motion. We define the optical flow vector on the image. As the optical flow is estimated from two consecutive images, it appears as a ݒԦ = ሺݑ,ݒሻ, where u and v represent the x and y components of the optical flow vector at a point displacement vector ݀Ԧ = ሺ݀௫,݀௬ሻ [6]. Where ݀௫ and ݀௬ denote the x and y components respectively of the displacement vector at a point. The Horn and Schunk algorithm [5] implemented in this work is based on differential solving schemes. It computes image velocity from numerical evaluation of spatiotemporal derivatives of image intensities. Let denote image intensity (or brightness) at the point (x,y) in the image plane at time t by E(x,y,t). Thus, the image domain is consequently assumed to be differentiable in space and time. The basic assumption in measuring image motion is that the brightness of a particular point in the time-varying image is constant, as a result, ܧሺݔ,ݕ,ݐሻ = ܧሺݔ + ߜݔ,ݕ +ߜݕ,ݐ +ߜݐሻ = 0, (2) 35 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 where ߜݔ and ߜݕ are the displacement in x direction and y direction, respectively, after a short interval of time ߜݐ. This assumption brings to the following condition, known as the “Optical Flow Constraint Equation”. If brightness varies smoothly with x, y, and t, Equation (2) can be expanded in a Taylor series and obtain ܧሺݔ,ݕ,ݐሻ+ ߜݔ డா డ௫ +ߜݕ డா డ௬ + ߜݐ డா డ௧ + ݁ = ܧሺݔ,ݕ,ݐሻ, (3) where e contains second and higher-order terms in ߜݔ, ߜݕ, and ߜݐ. Subtracting ܧሺݔ,ݕ,ݐሻ on both sides, dividing by ߜݐ, and taking the limit as ߜݐ → 0, we have ݀ݔ ݀ݐ +߲ܧ ߲ܧ ݀ݐ +߲ܧ ݀ݕ ߲ݐ = 0 ߲ݕ ߲ݔ ௗ௧ = ݒ, and using the abbreviations ܧ௫ = డா ௗ௧ = ݑ and ௗ௬ where ௗ௫ ܧ௫ݑ +ܧ௬ݒ + ܧ௧ = 0. finally obtain Equation (5), which is called the Optical Flow Constraint Equation (4) డ௫, ܧ௬ = డா డ௬, and ܧ௧ = డா డ௧, we (5) The components u and v are used to approximate the velocity component in the direction of the spatial gradient of the image intensity. Therefore, this is an under-constrained equation, since only the motion component in the direction of the local gradient of the image intensity function may be estimated, this is known as “the aperture problem”. It can be seen rewriting Equation (5) in the form, ሺܧ௫,ܧ௬ሻ∙ሺݑ,ݒሻ = −ܧ௧, (6) thus the component of the movement in the direction of the brightness gradient ሺܧ௫,ܧ௬ሻ is − ா೟ ටாೣమାா೤మ . (7) From Equation (7), it is seen that we cannot determine the component of the optical flow at right angles to the brightness gradient. As a result, the flow velocity (u,v) cannot be computed locally without introducing additional constraints [5]. Horn and Schunck introduced a smoothness constraint, supposing that the motion field is smooth over the entire image domain [5]. This constraint, measures the departure from smoothness in the velocity flow. To express the smoothness constraint is required to minimize the square of the magnitude of the gradient of the optical flow velocity as follows, ݁௦ = ඵሺሺݑ௫ଶ +ݑ௬ଶሻ+ሺݒ௫ଶ +ݒ௬ଶሻሻ݀ݔ݀ݕ the error in the optical flow constraint equation can be expressed as, ݁௘ = ඵሺܧ௫ݑ + ܧ௬ݒ +ܧ௧ሻଶ݀ݔ݀ݕ (8) (9) Horn and Schunck’s algorithm computes an estimation of the velocity field (u, v) that minimizes the sum of the errors (Equation 9) for the rate of change of image brightness in equation (3), and the measure of the departure from smoothness in the velocity flow, equation (8) [5]. By combining equation (3) and (8) with a suitable weighting factor, and by using calculus of variation to minimize the total error, it is formulated an iterative solution for optical flow: ݑ௡ାଵ = ݑത௡ − ܧ௫൫ܧ௫ݑത௡ + ܧ௬ݒı௡ + ܧ௧൯ ∝ଶ+ܧ௫ଶ + ܧ௬ଶ (10) 36 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 ݒ௡ାଵ = ݒı௡ −ܧ௬൫ܧ௫ݑത௡ +ܧ௬ݒı௡ + ܧ௧൯ Where superscripts indicate the iteration number, subscripts refer to derivation, and α is a positive constant know as smoothness factor. ଶ ଶ ൅ ܧ௬ Ýଶ൅ ܧ௫ (11) 3. EXPERIMENTAL RESULTS 3.1. Dynamic NIRS response to acupuncture stimulation The hemodynamic response was visualised by computing the colour representation in both probes. Ten serial samples from subject 4 based on HbO data from the 24 channels are presented in two separated images. Left image represents channels 13 to 24, while right image represents channels 1 to 12; the ten samples were taken every 22 seconds. Figure 5B shows the hemodynamic response in channel 1 while Figure 5C shows the hemodynamic response in channel 17. Red concentrations in Figure 5A indicate higher hemodynamic response to acupuncture stimulation. The increase and decrease of HbR was less significant than the response of HbO2, this effect was observed on both sides of the head. In addition, all patients showed a similar response pattern however different lags and magnitudes among channels were obtained for each subject. Hemodynamic response to acupuncture stimulation is explained as follows. 1 A 2 8 0 3 0 1 5 2 0 7 8 0 3 0 1 5 2 0 1 3 1 8 2 3 1 3 1 8 2 3 1 6 2 1 1 6 2 1 1 4 1 9 2 4 1 4 1 9 2 4 1 7 2 2 1 7 2 2 0 1 5 0 0 1 5 0 2 1 7 0 2 0 2 1 7 0 2 0 9 0 4 0 9 0 4 0 1 1 6 0 1 0 1 1 6 0 1 0 3 1 5 2 0 8 1 5 2 0 1 3 1 8 2 3 1 3 1 8 2 3 1 6 2 1 1 6 2 1 1 4 1 9 2 4 1 4 1 9 2 4 1 7 2 2 1 7 2 2 0 1 5 0 0 1 5 0 2 1 7 0 2 0 2 1 7 0 2 0 9 0 4 0 9 0 4 0 1 1 6 0 1 0 1 1 6 0 1 0 8 0 3 0 8 0 3 0 4 0 1 5 0 2 1 7 0 2 0 9 0 4 0 1 1 6 0 1 0 8 0 3 0 1 5 2 0 1 3 1 8 2 3 9 0 1 5 0 2 1 7 0 2 0 9 0 4 0 1 1 6 0 1 0 8 0 3 0 1 5 2 0 1 3 1 8 2 3 1 6 2 1 1 6 2 1 1 4 1 9 2 4 1 4 1 9 2 4 1 7 2 2 1 7 2 2 5 1 5 2 0 1 3 1 8 2 3 10 1 5 2 0 1 3 1 8 2 3 8 0 3 0 8 0 3 0 1 6 2 1 1 6 2 1 1 4 1 9 2 4 1 4 1 9 2 4 1 7 2 2 1 7 2 2 0 1 5 0 0 1 5 0 2 1 7 0 2 0 2 1 7 0 2 0 9 0 4 0 9 0 4 0 1 1 6 0 1 0 1 1 6 0 1 0 1 3 1 8 2 3 1 3 1 8 2 3 1 6 2 1 1 6 2 1 1 4 1 9 2 4 1 4 1 9 2 4 1 7 2 2 1 7 2 2 0 1 5 0 0 1 5 0 2 1 7 0 2 0 2 1 7 0 2 0 9 0 4 0 9 0 4 0 1 1 6 0 1 0 1 1 6 0 1 0 8 0 3 0 8 0 3 0 C 1 5 2 0 6 1 5 2 0 0.5 0.4 0.3 0.2 0.1 0 B ) m m M m * i l ( n b o g o m e H 1 2 3 4 5 6 7 8 9 10 ) m m * M m ( -0.1 -0.2 220 440 660 880 1100 1320 1540 1760 1980 2200 Samples (10/sec) Oxy DeOxy Total 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 1 2 3 4 5 6 7 8 9 10 220 440 660 880 1100 1320 Samples (10/sec) 1540 1760 1980 2200 Oxy DeOxy Total i l n b o g o m e H Figure 5. Dynamic haemoglobin response during experiment. (A) Ten sequential NIRS images, obtained every 22 seconds, and numbered from 1 to 10. The images show the dynamic change of HbO trough 24 channels. The two images represent each side of the cerebral hemisphere. (B) The dynamic trace recorded from channel 1 on the right side of the head. The number along upper Y axis corresponds to the numbered images in A. 37 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 (C) The dynamic trace recorded from channel 17 on the left side of the head. The number along upper Y axis corresponds to the numbered images in A. Figure 6 exhibits the pain dynamic changes of haemoglobin (Hb) of Oxygenated Hb (HbO, in blue), Deoxygenated Hb (HbR, in red), and Total Hb (HbT, in green); For illustration purposes, data was taken from channel 3. Markers (dot lines) show the start and end of needle insertion stimulation (task 1), with a duration of 6 seconds. HbT is the combination of HbO and HbR, in other words is the sum of HbO and HbR. In a typical response after each acupuncture stimulation, the HbO presented a slight drop followed immediately for a constant rise to reach a peak, then a gradual decrease to resting level. On the other hand, HbR presented a slight increase then a moderate decrease to reach its lowest point; this variation was contrasting the reaction of HbO. In addition, HbT was dependent of both HbO and HbR, therefore it varied according to the behaviour of these two signals. It is to note that the estimations of HbO are more accurate than HbR and HbT [7, 8], for this reason HbO was used to present and analyse the data from this section onwards. ) m m * M m ( i l n b o g o m e H 0.4 0.3 0.2 0.1 0 -0.1 -0.2 Oxy DeOxy Total B C A 300 350 400 450 500 550 600 Samples (10/sec) Figure 6. Distinctive dynamic response obtained after acupuncture stimulation. A) small drop of HbO after stimulation, B) followed by a constant increase until reach a peak, C) then a gradual decrease to resting level. 3.2. Pain responses of HbO to Acupuncture Stimulation Needle insertion (task 1) was the first acupuncture manipulation in the experiment. It was applied at thirty seconds after commencement of experiment and lasted for six seconds. Each acupuncture manipulation was applied within the time indicated by the dot lines (markers) in Figure 7, four channels were used only for illustration purposes. Once the insertion was done (within samples 301 and 361), the needle was twirled until de-qi (Chinese for obtaining the qi or arriving at the qi) was reached. Once the patient manifested the numbness/heaviness in the arm, the acupuncturist stopped the stimulation. This de-qi sensation was approximately reached three seconds after needle insertion. Figure 7 shows the acupuncture pain response after needle insertion (T1). Immediately after reaching the highest pain response, approximately after 14 seconds, HbO dropped towards baseline. It is worth mentioning that in all subjects the highest peak was exhibited after this first stimulus (needle insertion). This drop-rise-drop behaviour was exhibited on both sides of the head (Figure 5B,C). Needle insertion was followed by a resting time of 30 seconds. This time was used for the patient to rest after the stimulus and let the haemoglobin signals go back to base line. It is worth mentioning that there is delay in reaction time between the physical stimulation and the NIRS signal among all the stimulations, and this is consistent with previous research using NIRS [9, 10]. 38 ) m m * M m i l ( n b o g o m e H y x O The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 T1 T2 T2 T2 T3 301 361 661 761 1161 1061 Samples (10/sec) 1461 1561 Ch12 Ch1 Ch7 Ch10 1861 1911 Figure 7. Pain Response after needle insertion. Needle twirl (task 2) by the acupuncturist and it aims to increase the de-qi sensation. The time frame for this stimulation is ten seconds (samples 661-761) and was carried out immediately after resting time (Rt1), refer to Figure 6. The Acupuncturist stopped the twirl manipulation early when the patient expressed full numbness before the ten seconds. Similarly, the signal pain behaviour exhibited a small drop first before increasing constantly for 8 seconds. At approximately 14 seconds after the start of the twirl stimulation a large peak of HbO was obtained. Again, this signal dropped towards baseline after reaching its highest peak. After the stimulus application, the patient was given 30 seconds to rest. The second (within samples 1061-1161) and third (within samples 1461-1561) twirl stimulation exhibited a different pattern than the first stimulation. These responses were smaller than the first response (see Figure 6), exhibiting a small increase and decrease of HbO after the stimulation. In contrast to the first needle twirl, the peak was reached after only three seconds in the second needle twirl; while for the last twirl, the stimulation response was slightly bigger than the previous twirl. The fact that the hemodynamic response decreased (compared to the first twirl stimulation) can be described by two factors: 1) the patient was familiar with the twirl manipulation and became adapted to this sensation, 2) by applying pressure in the hogu point (well known by its analgesic properties) the patient experienced a painkiller effect that decrease the effect of pain by needle manipulation. After the stimulus application, the patient was given 30 seconds to rest. Hemodinamics of needle removal (task 3) are similar to task 1, in this stage the removal was done within a time frame of five seconds (samples 1861-1911). This stimulation was carried out immediately after the fourth resting time (Rt4). In this manipulation, the acupuncturist removed the needle with a fast abrupt pull. Similarly, to all previous stimulations, the pain response to needle manipulation exhibited a small drop of HbO followed by an increase to reach a peak in approximately four seconds after needle removal. After needle removal patient was given 40 seconds to rest which was used to let the signals go back to base line and to end the experiment. 3.3. Dominant pattern analysis After visual analysis, lags between channels were observed which suggested dominant channels on both sides of the head. As shown in Figure 8, images taken after acupuncture stimulation show that around channel 7 (dotted lines) the magnitude of HbO is stronger and appears to be the centre of radiation towards surrounding channels. This is a strong indication that the area near channel 7 39 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 (Ch7) is a dominant area for this particular patient. Figure 8 also reveals that surrounding channels (Ch4, Ch5, Ch9 and Ch10) had a stronger relationship with Ch7 since their magnitude increased similarly. On the other hand, channels Ch1, Ch6 and Ch11 had a week relationship since their magnitude increased later at a slower rate. 05 10 05 10 05 10 02 07 12 02 07 12 02 07 12 04 09 04 09 04 09 01 06 11 01 06 11 01 06 11 03 08 03 08 03 08 Figure 8. Images taken every five seconds after acupuncture stimulation showing dominant channel. Based on the assumption that there is a relationship among channels, the cross correlation analysis was computed. Figure 9 shows the time-dependant cross correlation between Ch7 and other channels for needle insertion stimulus. The results show that channel 7 and surrounding channels have a small lag time that confirms the results of the visual analysis. An example of this is the lag between Ch7 and Ch9 (τ=+0.2 sec), or Ch7 and Ch6 (τ=+2.5 sec); these results reflect the lag time between this dominant region and the time delay among peripheral channels. In addition, no delay was observed between Ch7 and surrounding channels Ch5 and Ch10. These results suggest that the hemodynamic of stimulation signals have a progressive movement from dominant regions towards peripheral regions. It is also important to note that following stimulation tasks showed a similar pattern, however the lag time is less significant, it suggests that lag/led time between channels is directly related to the stimulation intensity. 7 h C s v n o i t a e r r o c s s o r C l 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 Stimulus 1 and rest time Ch1 (lag=1.0sec) Ch2 (lag=0.1sec) Ch4 (lag=0.0sec) Ch5 (lag=0.0sec) Ch6 (lag=2.5sec) Ch9 (lag=0.2sec) Ch10 (lag=0.0sec) Ch12 (lag=0.1sec) 5 10 15 20 25 30 35 40 Lag time (sec) Figure 9. Cross correlation comparison between dominant channel and surrounding channels. 3.4. Direction of Pain Signal Information. Following the results obtained by cross correlation analysis, optical flow was used to determine the direction of NIRS signals through the channels. Figure 10 shows the optical field computed with two consecutive images from channels 1-12 after needle insertion of patient four. It can be seen that flow of NIRS data goes from channel 7 and it radiates towards surrounding channels. It supports the previous finding using cross correlation analysis about the existence of dominant 40 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 channels. This “dominant area” or “dominant channel” behaviour was found on both cerebral hemispheres and is present in all patients. 05 10 05 10 02 07 12 02 07 12 04 09 04 09 01 06 11 01 06 11 03 08 03 08 30 25 20 15 10 5 0 0 5 10 15 20 25 30 Figure 10. Optic flow computed from two consecutive images (samples 300 and 301) of patient four. It is also important to note that based on optic flow analysis, pain relationship can be made. Figure 11 presents optic flow fields of four pair of samples taken from the complete acupuncture manipulation process. In this image, only four channels (Ch2, Ch7, Ch10, and Ch12) are showed in the image for demonstration purposes. Figure 11A shows the optic flow at four different stages through the experiment, the first two OF images exhibit the radiation pattern from Ch7 to peripheral channels; this outer movement is directly related to the slope increase of the pain data. On the other hand, the last two OF images represent an internal movement from peripheral channels to Ch7; this inner movement is related to the slope decrease in pain stimulation. In other words, it is possible to link the OF result with the intensity and progression of pain through the experiment. 30 25 20 15 10 5 0 0 1 A B 2 30 25 20 15 10 5 0 0 5 10 15 20 25 30 3 30 25 20 15 10 5 0 0 5 10 15 20 25 30 4 30 25 20 15 10 5 0 0 5 10 15 20 25 30 5 10 15 20 25 30 ) * m m M m ( i l n b o g o m e H y x O 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 upward slopes downward slopes 301 361 661 761 1161 1061 Samples (10/sec) 1461 1561 Ch12 Ch1 Ch7 Ch10 1861 1911 Figure 11. Optic flow results from four consecutive pair of samples, note the slope direction where the sample where taken. A) Optic flow fields showing the outer and inner flow of information between Ch7 and peripheral channels. B) NIRS data show the slope direction that reflects the intensity level of pain stimulation. 41 The International Journal of Multimedia & Its Applications (IJMA) Vol.7, No.2, April 2015 4. CONCLUSIONS In this paper, the distinctive hemodynamic curve response to three types of acupuncture stimulation was presented. It was showed that the typical pain signal presents a small drop first to continue with a constant rise until it reaches a maximum to drop again towards resting level. This behaviour was consistent in all patients after the three types of acupuncture manipulations: needle insertion, needle twirl, needle removal. We have also implemented cross correlation analysis and optical flow analysis within the context of diffuse optical imaging of acupuncture stimulation. Cross correlation confirmed visual analysis by identifying time-dependant relationship among channels. Regions where the neural activity is dominant, such as Ch7 in patient four, were also identified. This dominant neural activity was present in all patients, however magnitude levels varied among patients. Optical flow corroborated pattern flow from dominant channels to peripheral channels. This method also distinguished the relationship between the pain stimulation and direction of pattern flow. It was found that an increase of pain stimulation was reflected as flow of NIRS information moving towards peripheral channels, while decrease of pain stimulation was reflected as flow of NIRS signals moving from peripheral channels to dominant channel. REFERENCES [1] E. R. Perl, "Ideas about pain, a historical view," Nature Reviews Neuroscience, vol. 8, pp. 71-80, 2007. [2] M. Bartocci, L. L. Bergqvist, H. Lagercrantz, and K. Anand, "Pain activates cortical areas in the preterm newborn brain," Pain, vol. 122, pp. 109-117, 2006. [3] C.-H. Lee, T. Sugiyama, A. Kataoka, A. Kudo, F. Fujino, Y.-W. Chen, et al., "Analysis for distinctive activation patterns of pain and itchy in the human brain cortex measured using near infrared spectroscopy (NIRS)," PloS one, vol. 8, p. e75360, 2013. [4] M. Verner, M. J. Herrmann, S. J. Troche, C. M. Roebers, and T. H. Rammsayer, "Cortical oxygen consumption in mental arithmetic as a function of task difficulty: a near-infrared spectroscopy approach," Frontiers in human neuroscience, vol. 7, 2013. [5] B. K. Horn and B. G. Schunck, "Determining optical flow," in 1981 Technical Symposium East, 1981, pp. 319-331. [6] B. D. Lucas and T. Kanade, "An iterative image registration technique with an application to stereo vision," in IJCAI, 1981, pp. 674-679. [7] Y. Otsuka, E. Nakato, S. Kanazawa, M. K. Yamaguchi, S. Watanabe, and R. Kakigi, "Neural activation to upright and inverted faces in infants measured by near infrared spectroscopy," NeuroImage, vol. 34, pp. 399-406, 2007. [8] M. Peña, A. Maki, D. KovacIJ ić, G. Dehaene-Lambertz, H. Koizumi, F. Bouquet, et al., "Sounds and silence: an optical topography study of language recognition at birth," Proceedings of the National Academy of Sciences, vol. 100, pp. 11702-11705, 2003. [9] Y. Honda, E. Nakato, Y. Otsuka, S. Kanazawa, S. Kojima, M. K. Yamaguchi, et al., "How do infants perceive scrambled face?: A near-infrared spectroscopic study," Brain research, vol. 1308, pp. 137- 146, 2010. [10] S. Boden, H. Obrig, C. Köhncke, H. Benav, S. Koch, and J. Steinbrink, "The oxygenation response to functional stimulation: is there a physiological meaning to the lag between parameters?," Neuroimage, vol. 36, pp. 100-107, 2007. 42
1502.02764
1
1502
2015-02-10T03:07:56
The Modeling and Quantification of Rhythmic to Non-rhythmic Phenomenon in Electrocardiography during Anesthesia
[ "q-bio.NC", "cs.CE" ]
Variations of instantaneous heart rate appears regularly oscillatory in deeper levels of anesthesia and less regular in lighter levels of anesthesia. It is impossible to observe this "rhythmic-to-non-rhythmic" phenomenon from raw electrocardiography waveform in current standard anesthesia monitors. To explore the possible clinical value, I proposed the adaptive harmonic model, which fits the descriptive property in physiology, and provides adequate mathematical conditions for the quantification. Based on the adaptive harmonic model, multitaper Synchrosqueezing transform was used to provide time-varying power spectrum, which facilitates to compute the quantitative index: "Non-rhythmic-to-Rhythmic Ratio" index (NRR index). I then used a clinical database to analyze the behavior of NRR index and compare it with other standard indices of anesthetic depth. The positive statistical results suggest that NRR index provides addition clinical information regarding motor reaction, which aligns with current standard tools. Furthermore, the ability to indicates the noxious stimulation is an additional finding. Lastly, I have proposed an real-time interpolation scheme to contribute my study further as a clinical application.
q-bio.NC
q-bio
Graduate Institute of Biomedical Electronics and Bioinformatics College of Electrical Engineering and Computer Science National Taiwan University Doctoral Dissertation The Modeling and Quantification of Rhythmic to Non-rhythmic Phenomenon in Electrocardiography during Anesthesia Author : Yu-Ting Lin Advisor : Jenho Tsao, Ph.D. February 2015 5 1 0 2 b e F 0 1 ] . C N o i b - q [ 1 v 4 6 7 2 0 . 2 0 5 1 : v i X r a "All composite things are not constant. Work hard to gain your own enlightenment." Siddh¯artha Gautama Abstract Variations of instantaneous heart rate appears regularly oscillatory in deeper levels of anesthesia and less regular in lighter levels of anesthesia. It is impossible to observe this "rhythmic-to-non-rhythmic" phenomenon from raw electrocardiography waveform in cur- rent standard anesthesia monitors. To explore the possible clinical value, I proposed the adaptive harmonic model, which fits the descriptive property in physiology, and provides adequate mathematical conditions for the quantification. Based on the adaptive har- monic model, multitaper Synchrosqueezing transform was used to provide time-varying power spectrum, which facilitates to compute the quantitative index: "Non-rhythmic- to-Rhythmic Ratio" index (NRR index). I then used a clinical database to analyze the behavior of NRR index and compare it with other standard indices of anesthetic depth. The positive statistical results suggest that NRR index provides addition clinical infor- mation regarding motor reaction, which aligns with current standard tools. Furthermore, the ability to indicates the noxious stimulation is an additional finding. Lastly, I have proposed an real-time interpolation scheme to contribute my study further as a clinical application. Keywords: instantaneous heart rate; rhythmic-to-non-rhythmic; Synchrosqueezing trans- form; time-frequency analysis; time-varying power spectrum; depth of anesthesia; electro- cardiography Acknowledgements First of all, I would like to thank Professor Jenho Tsao for all thoughtful lessons, discus- sions, and guidance he has provided me in the last five years. His immense knowledge in electrical engineering has also been a great addition to assisting me in my research. I thank Professor Hsiao-Lung Chan, Professor Kung-Bin Sung, Professor Nien-Tsu Huang, and Professor Hau-tieng Wu for joining my dissertation defense committee. I would also like to thank Professor Chih-Ting Lin for his great advice in my proposal defense. I would like to thank Professor Yi Chang and all my colleagues in the department of anesthesia, Shin Kong Wu Ho-Su Memorial Hospital, for supporting me on seeking the answer to my clinical question. I am particularly grateful to Cheng-Hsi Chang M.D. for his constant assistance and advice in regards to my clinical practice and my research. I would like to thank Professor Hau-tieng Wu for his inspiration and collaboration. It has been a great blessing that we were both able to work together and understand the problems from the perspective as a scientist, engineer, and physician. This has helped us drawn wonderful and meaningful results. I am very grateful for the valuable assistance and discussion from all the members in Signal Processing Laboratory. I am particularly grateful to Pei-Feng Lin, Ming-huang Chen, Zongmin Lin for all their valuable assistance and inspiring discussions during my stay at the NTU campus. I would like to express my sincere gratitude to Hui-Hsun Huang M.D. in the department of anesthesia, National Taiwan University Hospital. His pioneer work on using time-frequency analysis to attack the IHR problem in anesthesia hugely inspired me since my residency. I would also like to thank Professor Charles K. Chui for his guidance in seeking a solution on real-time interpolation. My acknowledgements would be incomplete without recognition of Shu-Shya Hseu M.D. and Huey-Wen Yien M.D. for their extraordinary mentorship since I was a medical student. My greatest appreciation are reserved for my family. My parents and brother have been the constant support of my life. My daughter and my son have always brought me joy. Lastly, I am most grateful to my wife's consistent dedication to me since we have met. iii Contents Abstract Acknowledgements Contents List of Figures List of Tables Abbreviations 1 Introduction 1.1 Field of Project . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 An Obscure Phenomenon . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Quantification for Depth of Anesthesia . . . . . . . . . . . . . . . . . . . . . 1.4 Project Motivation and Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5 Thesis Outline 2 Background and Previous Work 2.1 Historical Background in Anesthesia . . . . . . . . . . . . . . . . . . . . . . 2.2 Current View on Anesthesia . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Perspective from Anesthesiology . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Anesthesia and Long-Term Mortality . . . . . . . . . . . . . . . . . . . . . . 2.5 Perspective from Physiology . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6 Physiology of Amplitude and Frequency Modulation . . . . . . . . . . . . . 2.7 Instantaneous Heart Rate and Heart Rate Variability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Integration of Multidisciplinary Backgrounds 2.8 ii iii iv vii viii ix 1 1 1 1 2 3 4 4 4 5 6 6 7 8 9 3 Theory and Modeling 10 3.1 Proposal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 3.2 Description of Proposed Methodology . . . . . . . . . . . . . . . . . . . . . 10 3.3 Adaptive Harmonic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 3.4 Non-rhythmic to Rhythmic Ratio . . . . . . . . . . . . . . . . . . . . . . . . 12 3.5 Non-rhythmic Component vs. Stochastic Process . . . . . . . . . . . . . . . 15 3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 iv Contents v 4 Time-Frequency Analysis, Reassignment, and Synchrosqueezing 17 4.1 Time-Frequency Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 4.2 Reassignment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 4.3 Synchrosqueezing Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 4.4 Multitaper Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 4.5 Multitaper Synchrosqueezing Spectrogram . . . . . . . . . . . . . . . . . . . 23 4.6 Adaptive Filter and Parametric Estimation . . . . . . . . . . . . . . . . . . 24 . . . . . . . . . . . . . 26 4.7 Common Condition in Physiology and Mathematics 5 Performance Evaluation using Clinical Data 27 5.1 Clinical Database . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 5.2 Anesthesia Protocol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 5.3 Data Acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 5.4 Statistical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 5.5 Serial Prediction Probability Analysis in Predicting Anesthetic Events . . . 30 5.6 Correlations with Sevoflurane Concentration . . . . . . . . . . . . . . . . . . 31 5.7 Algorithm of Serial Prediction Probability (sPK) Analysis . . . . . . . . . . 31 6 Results from Clinical Database 33 6.1 Visual Information in TF Plane . . . . . . . . . . . . . . . . . . . . . . . . . 33 6.2 Multiple Component Phenomenon . . . . . . . . . . . . . . . . . . . . . . . 34 6.3 Ability to Predict Anesthetic Events . . . . . . . . . . . . . . . . . . . . . . 37 6.4 Correlation with Sevoflurane Concentration . . . . . . . . . . . . . . . . . . 41 6.5 Quantitative Results of Noxious Stimulation . . . . . . . . . . . . . . . . . . 42 6.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 7 Implications to Anesthesiology and Physiology 49 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 7.1 Main Findings 7.2 Genuineness of NRR index . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 7.3 Existing Findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 7.4 Physiologic Interpretation of NRR . . . . . . . . . . . . . . . . . . . . . . . 50 7.5 Clinical Application of the NRR Index . . . . . . . . . . . . . . . . . . . . . 52 7.6 Potential Index in Noxious Stimulation . . . . . . . . . . . . . . . . . . . . . 52 7.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 8 Real-time Processing Using the Blending Operator 54 8.1 Real-time Processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 8.2 Technique Obstacles for Real-time Processing . . . . . . . . . . . . . . . . . 55 8.3 B-spline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 8.4 Quasi-interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 8.5 Local Interpolation Operator . . . . . . . . . . . . . . . . . . . . . . . . . . 59 8.6 Blending Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 8.7 Real-time tvPS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 9 Conclusion 63 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 9.1 Research Findings 9.2 Accomplishments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 9.3 Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 Contents vi 9.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 Bibliography 66 List of Figures 1.1 Rhythmic-to-Non-rhythmic phenomenon . . . . . . . . . . . . . . . . . . . . 2.1 R-R peak interval and instantaneous heart rate . . . . . . . . . . . . . . . . 2 8 3.1 The Calculation of Rhythmic Power and Non-rhythmic Power . . . . . . . . 14 4.1 Color Graph: STFT spectrogram vs. Reassigned spectrogram . . . . . . . . 19 4.2 STFT spectrogram vs. Reassigned spectrogram . . . . . . . . . . . . . . . . 21 4.3 The time-varying power spectrum: multitaper Synchrosqueezing spectrogram 24 4.4 Demonstration of multitaper Synchrosqueezing using artificial data . . . . . 25 6.1 Dynamic Feature in Anesthesia . . . . . . . . . . . . . . . . . . . . . . . . . 34 6.2 Rhythmic-to-Non-rhythmic phenomenon, NRR index and depth of anesthe- sia varying during controlled ventilation . . . . . . . . . . . . . . . . . . . . 35 6.3 Rhythmic-to-Non-rhythmic phenomenon, NRR index and depth of anesthe- sia during spontaneous breathing . . . . . . . . . . . . . . . . . . . . . . . . 36 6.4 Skin Incision during Anesthesia . . . . . . . . . . . . . . . . . . . . . . . . . 37 6.5 Rhythmic-to-Non-rhythmic Phenomenon Demo no.1 . . . . . . . . . . . . . 38 6.6 Rhythmic-to-Non-rhythmic phenomenon Demo no.2 . . . . . . . . . . . . . 38 6.7 Rhythmic-to-Non-rhythmic phenomenon Demo no.3 . . . . . . . . . . . . . 39 6.8 Multiple Components in IHR during Anesthesia . . . . . . . . . . . . . . . . 39 6.9 Two Components in IHR during Anesthesia . . . . . . . . . . . . . . . . . . 40 6.10 Multiple components and rhythmic-to-non-rhythmic phenomenon . . . . . . 40 6.11 Multiple components and rhythmic-to-non-rhythmic phenomenon no.2 . . . 41 6.12 Serial PK Analysis: Loss of Consciousness . . . . . . . . . . . . . . . . . . . 42 6.13 Serial PK Analysis: Skin Incision . . . . . . . . . . . . . . . . . . . . . . . . 43 6.14 Serial PK Analysis: First Reaction . . . . . . . . . . . . . . . . . . . . . . . 44 6.15 Serial PK Analysis: Return of Consciousness . . . . . . . . . . . . . . . . . . 45 6.16 Serial PK Analysis: Endotracheal Intubation . . . . . . . . . . . . . . . . . 46 9.1 Diagram of "rhythmic-to-non-rhythmic phenomenon" . . . . . . . . . . . . 64 vii List of Tables 4.1 Comparison of TF analysis methods . . . . . . . . . . . . . . . . . . . . . . 26 6.1 Representative Value of Indices . . . . . . . . . . . . . . . . . . . . . . . . . 47 6.2 Correlation to Estimated Sevoflurane Effect-site Concentration . . . . . . . 48 viii Abbreviations NRR HR IHR HRV ECG Non-Rhythmic-to-Rhythmic Heart Rate Instantaneous Heart Rate Heart Rate Variability ElectroCardioGraphy PK analysis Prediction probability analysis BIS index Bispectral index PS tvPS STFT CWT TF RRI EDR Power Spectrum time-varying Power Spectrum Short Time Fourier Transform Continuous Wavelet Transform Time-Frequency R-peak to R-peak Intevral ECG-Derived Respiration HF power High Frequency power LF power Low Frequency power tvHF power time-varying High Frequency power tvLF power time-varying Low Frequency power ix To My Family and to the Memory of My Grandparents. x Chapter 1 Introduction 1.1 Field of Project The human body undergoes unique conditions when under anesthesia. Anesthesiologists administer anesthetics, also known as anesthetic medicine, dependent on the patient's physiological information. The physiological information may include the patient's heart rate, blood pressure, and electroencephalography (EEG) as this information reflects the inner dynamics of the human body. My study will present a new perspective on the physiological information that could contribute to the clinical practice of anesthesia. 1.2 An Obscure Phenomenon As a practicing anesthesiologist, I have noticed that the instantaneous heart rate (IHR) fluctuates dependent on the various levels of anesthesia -- it became regularly oscillatory in deeper level of anesthesia, whereas it became less regularly in lighter level.(see Fig.1.1) This phenomenon, I refer to as "rhythmic-to-non-rhythmic phenomenon", has been described qualitatively in the past. However, to the best of my knowledge, no quantitative method has been proposed. Also, the underlying physiological mechanism has not been addressed. 1.3 Quantification for Depth of Anesthesia This phenomenon has induced three questions -- 1) how to quantify this phenomenon into a useful index, 2) whether the quantitative index can reflect the level of anesthesia, and 3) what are the underlying physiological mechanisms. 1 Chapter 1. Introduction 2 Figure 1.1: The tracing of R-R peak interval (RRI) represents instantaneous heart rate (IHR) ex- hibiting "rhythmic" or "non-rhythmic" patterns during anesthesia. This phenomenon is difficult to be read from the raw electrocardiography (ECG) waveform. Interestingly, the ECG waveform is mandatory in standard anesthesia monitor, but IHR is not. For an anesthesiologist, information from this phenomenon is unavailable in the clini- cal perspective. From Fig.1.1, it is apparent that the "rhythmic-to-non-rhythmic" phe- nomenon can be seen from the tracing of instantaneous heart rate, but not from raw electrocardiographic (ECG) waveform. Although real-time monitoring of ECG is manda- tory in the modern anesthesia monitor, the tracing of instantaneous heart rate is not dis- played. Even though an anesthesiologist was aware of this phenomenon, he cannot obtain "rhythmic-to-non-rhythmic" phenomenon information from modern anesthesia monitor to improve the anesthetic management. If NRR index can provide additional clinical value, I believe that it could help in future developments of standard anesthesia monitor. 1.4 Project Motivation and Goals My clinical observation leads to an intuition that this phenomenon arises from the in- ner physiologic dynamics under surgery and anesthesia. Hence, the quantification could provide valuable information to help anesthesiologists in administering anesthetics. Pursuing these questions involves model construction based on knowledge of three distinct fields: signal processing, clinical anesthesiology and physiology. Then, I will propose a quantitative index, referred to as "non-rhythmic-to-rhythmic ratio"(NRR) to quantify the phenomenon. The computation of NRR index requires multitaper Synchrosqueezing transform, a recently developed time-frequency analysis technique. Chapter 1. Introduction 3 I used a clinical database to explore the property of NRR index. Meanwhile, NRR index was compared with standard depth-of-anesthesia indices used in current clinical practicing to understand its performance. 1.5 Thesis Outline My thesis will begin with a literature review of the clinical anesthesiology, neural physiol- ogy, and time-frequency analysis. Bringing multidisciplinary knowledge up to date leads to a theory for the possible mechanism of rhythmic-to-non-rhythmic phenomenon. I proposed a model as the foundation of further analysis using time-frequency analysis. The adaptive harmonic model is compatible in both ways: fitting the descriptive property in physiol- ogy, and providing adequate mathematical conditions for the quantification. Based on the adaptive harmonic model and multitaper Synchrosqueezing transform, NRR index was developed to quantify rhythmic-to-non-rhythmic phenomenon. I used a clinical database to analyze the behavior of NRR index, and compare it with other current standard indices of anesthetic depth. The statistical result shows positive value in clinical: NRR index provides information regarding motor reaction earlier and better than all the other depth- of-anesthesia indices. The result suggests a possible anatomic area and functional part in the brain that mediate the rhythmic-to-non-rhythmic phenomenon. To benefit each individual patient undergoing surgery and anesthesia, NRR index should be implemented in real-time. An optimal spline interpolation technique is proposed for the development of "real-time NRR index". Chapter 2 Background and Previous Work 2.1 Historical Background in Anesthesia More than seventy years ago, Guedel made a remark on the connection between anesthesia and respiratory patterns[1]. It appears that there is a direct relation between anesthetic depth and respiratory patterns. A deeper level of anesthesia results in regular respiratory patterns while a lighter level of anesthesia results in irregular respiratory patterns. Dur- ing that era, depth of anesthesia were concluded from physical signs, such as the changes in pupil size and movement of the eyeballs. However, the introduction of muscle relax- ant in clinical anesthesia has hindered the observation of muscle movement and Guedel's assessment method. Even though Guedel's observations became less known, this "rhythmic-to-non-rhythmic" phenomenon has been documented in other literature. About twenty years ago, Kato used power spectrum to analyze IHR in anesthesia[2]. His study showed that the spectrum appears to be concentrated during deeper level of anesthesia; whereas the spectrum is dispersing in lighter level of anesthesia. Kato also mentioned the respiratory effect on this phenomenon. It is possible that the power spectrum technique helps to capture the obscure phenomenon in IHR at that era. 2.2 Current View on Anesthesia Today, we would like to further explain this phenomenon through the advancement of various scientific fields in signal processing, anesthesiology, and neural physiology. Because of the advancement in anesthesiology, we now understand that anesthetics exert effect mainly on the brain. Distinct anatomic location mediates different functions, with 4 Chapter 2. Background 5 differential susceptibility to same anesthetics. Furthermore, each patient has his own unique genetic profile that renders diverse response to anesthetics. The inter-individual genetic variability, plus the dynamic influence of surgical procedure, make it difficult to determine appropriate anesthetic levels based on the dosage of anesthetics or single index of anesthetic depth. There is strong evidence that anesthetic management can improve the long-term health of patients, it is important to have monitoring instruments facilitating a comprehensive anesthetic management specific for each unique patient. Hence, the quantification of "rhythmic-to-nonrhythmic" phenomenon should be worthwhile. Current advancements in signal processing has resulted in the development of time-frequency analysis. In particular, reassignment and Synchrosqueezing technique were not readily ac- cessible in the past. Furthermore, current advancements in neural physiology has also provided more insight into the underlying mechanism of the "rhythmic-to-non-rhythmic" phenomenon, which has not been addressed according to my best knowledge. A more detail literature review regarding three main disciplines is as follows. 2.3 Perspective from Anesthesiology Anesthesia comprises several components, including hypnosis, analgesia, immobility, am- nesia, and autonomic nerve system stability[3 -- 6]. For each patient, the dosage for different anesthetics is tailored to the patient's specific surgical needs. Therefore, anesthesiologists ® need to continuously monitor the patient's response to the anesthesia. Bispectral Index (BIS), an index computed from electroencephalography(EEG), is the current standard tools used to monitor depth-of-anesthesia for either the hypnotic component or the level of sleepiness. However, analgesia component, commonly known as the inhibition of pain sensation in anesthesia, requires an indirect interpretation of physiologic information such as heart rate (HR) and blood pressure[7 -- 11]. It is impossible to achieve optimal anesthe- sia using only one kind of anesthetics. Also, anesthesiologists are unable to monitor the depth of anesthesia by using merely one index, consequently inadequately measuring anes- thetic effects. For a more comprehensive administration of anesthesia, it may be helpful to integrate different types of information to come to a decision. It is important to understand what is wrong with simply relying on BIS monitor. BIS is a processed EEG, which measures the activity of the frontal cortical area only. On the other hand, amnesia is the suppression of memory, which is governed by medial temporal cortical area. Analgesia is the suppression of pain sensation, which is governed by various subcortical areas. Immobility is the inability of movement which is typically controlled by the thalamus and spinal cord[7, 12, 13]. Lastly, the entire autonomic nerve system is controlled by the brainstem[14 -- 16]. Although BIS index is the current gold standard of Chapter 2. Background 6 anesthetic depth monitor, BIS sensor is unable to measure the activities from all above listed anatomical areas except the frontal cortex. Literature in recent years indicates that anesthetics exert differential effects on different brain areas[3, 7, 17]. Thus, there is a considerable risk measuring frontal cortex to conclude the anesthetic effect on other brain areas. For example, researchers have reported that the BIS monitor cannot reduce the rate of intra-operative awareness, which requires adequate memory function suppression[18]. 2.4 Anesthesia and Long-Term Mortality Anesthesia is more than merely falling asleep. Evidences support the connection between anesthesia and long-term mortality. Too deep of hypnotic index (BIS index) for too long in surgery is associated with worse one-year mortality[19]. It has also been reported that different anesthetics cause different immune responses, which affect tumor metastasis[20]. Evidence suggests incorporating "stress reduction" strategy into anesthetic management leads to longer life expectancy. First, infants who needed surgery for their congenital heart disease had lower long term mortality rate if they received adequate analgesics[21]. Second, more adult patients who underwent major surgeries survived one year later if they receive adequate medication to stabilize autonomic nerve system[22, 23]. Lastly, when epidural anesthesia, a regional anesthesia technique, is combined with general anesthesia for patients undergoing major surgeries, studies have showed that epidural anesthesia reduces stress and pain, leads to faster recovery, and overall better outcome[24]. There are more compelling studies that support the benefit of anesthesia management on long term outcome. These evidences suggest that a comprehensive optimization of anesthesia is worthwhile. In summary, anesthesia comprises several components. A comprehensive anesthesia man- agement that covers several components could benefit patient for many years. Liteurature survey also brings hope that quantitative analysis of "rhythmic-to-non-rhythmic" phe- nomenon could be useful to monitor the anesthetic effect on subcortical regions. 2.5 Perspective from Physiology From a physiological perspective, the underlying mechanism of the "rhythmic-to-non- rhythmic" phenomenon could be partially explained by breathing mechanism and car- diopulmonary coupling[14]. It is known that the neural respiratory control comprises of two systems, 1) the involuntary automatic control system and 2) the voluntary control system. The involuntary automatic control system is controlled by the respiratory center Chapter 2. Background 7 in the brainstem. The voluntary control system is controlled by the forebrain[25]. The anatomies of these two systems are distinct from one another. The respiratory neural signals from these two systems compete with each other but are both integrated at the spinal cord to control the respiratory motor neuron. During spontaneous respiration, the involuntary automatic respiratory pattern generated in the preBotzinger complex [25 -- 29], is rhythmic. On the contrary, the breathing pattern of the voluntary respiratory motor control is non-rhythmic, which involves larger cerebral areas, including cortical processing and thalamic integration[25, 30 -- 34]. The suppression effect of anesthetics on the forebrain, including the cortex and the tha- lamus, is stronger than that on the brainstem[35]. Meanwhile, studies have shown that the preBotzinger complex, the involuntary respiratory control center, is less susceptible to the anesthetics[36, 37]. These relations suggest that under deeper level of anesthesia, the non-rhythmic respiratory is more suppressed than the rhythmic respiratory center. Hence, the IHR exhibits oscillation that is more rhythmic during deeper anesthetic level. Contrarily, the non-rhythmic respiratory center, being less suppressed during lighter level of anesthesia, causes activity more non-rhythmic in IHR. This concludes that IHR exhibits more nonr-hythmic. In summary, literature in physiology supports the correlation that the differential anesthetic effects between the involuntary control system and voluntary control system induces the "rhythmic-to-non-rhythmic" phenomenon in IHR. Thus, I hypothesize that IHR exhibits the central respiratory activity via cardiopulmonary effect[14], and it reflects the integration of rhythmic and non-rhythmic respiratory activ- ities. Although more evidence is necessary to clarify this hypothesis, I propose using the NRR quantification methodology as a potential tool to evaluate the depth of anesthesia from a different aspect than using EEG-based monitoring. 2.6 Physiology of Amplitude and Frequency Modulation The human body constantly regulates the respiration to meet the requirement of metabolism. In other words, the human body has to inhale oxygen (O2), and expel carbon dioxide (CO2) adequately. The respiratory control center in the brainstem receives the feedback infor- mation of oxygen concentration and the acidity to modulates the volume and the rate of breath. The acidity is due to the fact that CO2 exists as hydrogen ion (H+) and carbonic acid in human body[16, 38]. Since CO2 are the metabolic products of human body, the H+ concentration and O2 concentration in the human body are gradually, but consistently, changing. Thus, we can describe that the respiration signal in the brainstem is oscillatory with amplitude Chapter 2. Background 8 modulation (AM) and frequency modulation (FM). The AM and FM are continuous and slow change in time. These physiologic features can be depicted in mathematical language for subsequent modeling and quantification in the next chapter. In summary, literature from physiology supports an integrated rhythmic and non-rhythmic neural activity reflecting the level of anesthesia. From a clinical anesthesiology perspective, technological advancements on monitoring various aspects of anesthetic depth can help improve the quality of overall patient care. 2.7 Instantaneous Heart Rate and Heart Rate Variability Figure 2.1: A schematic diagram for the R-R interval (RRI) signal derived from the raw lead III electrocardiographic (ECG) signal. We refer to RRI signal as instantaneous heart rate. The diagram also showed the ECG-derived respiration (EDR) The term heart rate is important to all anesthesiologists and is usually referred to as the number of heartbeats within one minute. To an anesthesiologist, an ongoing heart rate indicates the stability of the heart condition, or the stress responses to surgical stimulation. The heart rate is mainly controlled by both sympathetic nerve and parasympathetic nerve system from the brainstem to the heart[16]. Due to the dynamic nature of patient's under surgery and anesthesia, we do not feel comfortable with the heart rate information at an one-minute interval. Instead, a standard requirement in anesthesia practice includes a continuous real-time display of the heart rate. In this thesis, the "rhythmic-to-non- rhythmic" phenomenon is only observable at the instantaneous level. Conversely, the Chapter 2. Background 9 average heart rate of a normal routine excludes the information of "rhythmic-to-non- rhythmic" phenomenon. Both the Instantaneous heart rate (IHR) and ECG derived respiration (EDR),(Fig.2.1) can be derived from the raw ECG waveform[39 -- 41]. Although we can say that the "rhythmic- to-non-rhythmic" phenomenon in IHR is a kind of "generalized" heart rate variability (HRV), a medical guideline has set up the standard method of HRV analysis[42]. Un- fortunately, the standard analysis techniques stated in this guideline cannot reveal the information of the "rhythmic-to-non-rhythmic" phenomenon. In particular, the SDNNx, pNNx or the power spectrum technique measures information in a specified period, which inevitably neglects the dynamic and transient changes within this period. The guideline discusses but does not resolve the potential issue of stationarity in the data. This also supports my desire to rely on the current scientific advancements on signal processes to resolve the current medical shortcomings in my project. 2.8 Integration of Multidisciplinary Backgrounds Integration of multidisciplinary backgrounds provides partially theoretical background for the mechanism of "rhythmic-to-non-rhythmic" phenomenon. Below is a list of key facts from various disciplines: 1. In respiratory control, non-rhythmic activity originates from the forebrain, whereas rhythmic activity originates from the brainstem. 2. The forebrain has far more neurons and synapses when compared to the brainstem. 3. Neural structures with more synapses are more susceptible to anesthetics 4. Respiration have an coupling effect on IHR Fact 1 and 2 are knowledge in neurology and neural anatomy. Fact 3 is from anesthesiology, and fact 4 is physiology knowledge. Each single fact seems ordinary when considered in its own disciplines. However, when I combine them together, the theoretical ground is justified. Then, we can proceed to the modeling problem in the next chapter. Chapter 3 Theory and Modeling 3.1 Proposal The "rhythmic-to-non-rhythmic" phenomenon in IHR[2, 43] suggests that IHR during anesthesia consists of two components: one more rhythmic, less affected by anesthesia, and one more non-rhythmic, suppressed by deep anesthesia, and the strength of these two components vary according to the anesthesia depth. I hypothesize that quantifying the "rhythmic-to-non-rhythmic" phenomenon will reflect the level of anesthesia. In order to quantify this change in IHR, I will introduce a novel index which I referred to as Non-rhythmic to Rhythmic Ratio (NRR) index. 3.2 Description of Proposed Methodology My hypothesis will be tested through a four step method. I will capture the changes in rhythm through IHR using the adaptive harmonic model. Next, I will use the multitaper Synchrosqueezing transform to extract the dynamic behavior from the ECG signals. In order to quantify "rhythmic-to-non-rhythmic" phenomenon, I will use the NRR index. Lastly, I will use a clinical database to compare the NRR index with other anesthetic depth indices. 3.3 Adaptive Harmonic Model It is important to consider which appropriate model to utilize to adequately describe and quantify this rhythmic-to-non-rhythmic phenomenon. Specifically, it is important 10 Chapter 3. Modeling 11 to recall the high difficulty on describing the inner dynamic of the human body through mathematical formulas. This difficulty is due to the fact that the human body is a complex system creating non-linear and unpredictable interactions with the outer environment. The drastic influences of surgery and anesthetic medication complicates the difficulty further. It is fairly simple to ruin the capability of a model to describe the reality by imposing unrealistic conditions. That is the most challenging task currently at hand when we try to model the biological phenomenon of our study. In order to do so we have to be extremely cautious on what conditions we choose and impose on our study model. After careful observation from a time-varying spectrum, I propose a relatively conservative model to analyze the rhythmic part of IHR. As is shown in Fig.1.1, the definition of rhythmic implies that the oscillation exhibits time-varying frequency and time-varying amplitude. However, the modulations are in a relatively slow rate. This presents the notion that the relative predictable oscillatory activity can be perceived as a rhythmic movement. Hence, a more conservative phenomenological modeling, also known as adaptive harmonic model, will be proposed as follows. The adaptive model integrates with the Synchrosqueezing transform creating a whole new scheme starting from the theoretical grounds to the clinical phenomenon. We start from motivating the technical model we consider to analyze IHR(t). It is commonly observed in clinical settings that stronger anesthetic levels are associated with more regular patterns in the instantaneous heart rate IHR(t), which is caused by the cardiopulmonary coupling effect under anesthesia [2, 43]. Based on this major observation, our treatment of the IHR(t) signal will be purely phenomenological; that is, the parameters and indices we will derive from the IHR(t) signal will be based solely on these signals themselves, and not on explicit, quantitative models of the underlying mechanisms. Since we mainly extract IHR from ECG signal as RRI, we will consider the following phenomenological model for RRI(t): RRI(t) = T (t) + A(t) cos(2πφ(t)), (3.1) where T (t) > 0, A(t) > 0 and φ(cid:48)(t) > 0. We call T (t) the trend, which captures the "average heart rate" over a long period; we require the trend to be positive, but not to be constant; we allow it to vary in time, as long as the variations are small enough. We call the derivative φ(cid:48)(t) of the function φ(t) the instantaneous frequency (IF) of RRI(t); we require IF to be positive, but not to be constant; we allow it to vary in time, as long as the variations are slight from one period to the next, i.e. φ(cid:48)(cid:48)(t) ≤ φ(cid:48)(t) for all time t, where  is some small, pre-assigned positive number. Likewise, We call A(t) the amplitude modulation (AM) of R(t), which should be positive and smaller than T (t), but is allowed to vary slightly as well, i.e. A(cid:48)(t) ≤ φ(cid:48)(t) for all time t. In summary, we have the following Chapter 3. Modeling conditions for the model 3.1 [44, 45]: 0 < A(t) < T (t), φ(cid:48)(t) > 0, A(cid:48)(t) ≤ φ(cid:48)(t), φ(cid:48)(cid:48)(t) ≤ φ(cid:48)(t), T (cid:48)(t) ≤  for all t. 12 (3.2) (3.3) If a RRI signal satisfied the model (3.1), we call it rhythmic; otherwise we call it non- rhythmic. It is possible that a function can be composed of both the rhythmic and non- rhythmic components. This seeming complicated model is actually a direct generalization of the Fourier harmonic model [44, 45]. Close inspecting the waveform of IHR, we can see that the conditions on φ(cid:48)(cid:48)(t) and A(cid:48)(t) are reasonable. lation with "small enough" frequency modulation and amplitude modulation. Without In other words, the rhythmic RRI exhibits an oscil- adequate knowledge of underlying mechanism, the adaptive harmonic model fulfills the phenomenological behavior of rhythmicity feature. 3.4 Non-rhythmic to Rhythmic Ratio The amount of variance inside the RRI signal are known as Heart Rate Variability (HRV). HRV has been proven to be closely related to the autonomic activity [42]. The most commonly used tool to quantify HRV is the power spectrum (PS) [42]. As useful as the PS is, however, in this study the PS cannot capture the momentary dynamics in the RRI time series. This issue limits the application to human under highly dynamic situation -- in particular, the "rhythmic-to-non-rhythmic" phenomenon in anesthesia. The main technical focus of this paper is to resolve this limitation. In this paper I will quantify the rhythmic-to-non-rhythmic pattern of the RRI series by us- ing a novel index referred to as Non-rhythmic to Rhythmic Ratio (NRR). NRR is motivated by the clinical finding that deeper anesthetic levels are associated with the appearance of more regular oscillatory patterns in IHR. Thus, under deeper anesthesia, IHR oscillates more regularly. We call this regular oscillatory IHR "rhythmic", which appears as a sharp peak on the PS. On the other hand, when the subject is under lighter anesthesia, IHR varies irregularly, and we call this IHR "non-rhythmic". The non-rhythmic IHR exhibits irregular and random-like behavior that appears as a plateau on the PS. As a result from this observation, the RRI time series is composed of a non-rhythmic component and a rhythmic component with various ratios. Based on the above facts, we hypothesize that the time-varying ratio change of the non-rhythmic and rhythmic components may reflect the anesthetic effect on the brain. Chapter 3. Modeling 13 To quantify the ratio of the rhythmic component and non-rhythmic component, we will use a new signal processing tool to fully capture the momentary RRI time series change. We will replace the PS with the time-varying power spectrum (tvPS) by applying a recently developed time-frequency analysis technique [43, 46], which is referred to as multitaper Synchrosqueezing transform [44, 45, 47]. The tvPS is a non-parametric generalization of the PS providing instantaneous PS information. We will then be able to trace the momentary strength of the rhythmic component and non-rhythmic component through analyzing the tvPS. By using the tvPS we will further extend the definition of the classical frequency do- mains HRV parameters. Recall that traditionally, the high frequency (HF) power, the low frequency (LF) power, and the low frequency to high frequency power ratio (LHR), are determined from the PS of the RRI time series. Vagal activity mainly contributes to HF power, meanwhile, sympathetic activity influences both LF power and LHR [42, 48]. By using the tvPS of RRI, we are able to obtain time-varying HRV parameters, including the time-varying high frequency power (tvHF), the time-varying low frequency power (tvLF) and the time-varying low frequency to high frequency power ratio (tvLHR). All these pa- rameters are suitable to provide a better understanding of the dynamical situations of anesthesia. The main index NRR is defined as the ratio between momentary non-rhythmic power to momentary rhythmic power. In mathematical terms, we compute the rhythmic component power by identifying the location of peak on tvPS. The non-rhythmic power is then defined as tvHF subtracted by the momentary rhythmic power. Finally, NRR is defined as (cid:18) Non-rhythmic power (cid:19) Rhythmic power , NRR = log10 where the value of NRR is high with non-rhythmic RRI and the value of NRR is low with rhythmic RRI. As previously discussed, the dynamical behavior of RRI(t) is captured by tvPS. Please see Figure 3.1 In practical calculation, the non-rhythmic power is calculated as the rhythmic power sub- tracted from the high frequency power of RRI. The rhythmic component power is calcu- lated as the sum of the power near the IF of the rhythmic component within the range of high frequency on the tvPS. We define the IF of the rhythmic component, denoted as fr(t), as the time varying position of the maximal power within the high frequency range on the tvPS of RRI(t): fr(t) := argmax 0.1≤ξ≤1 MRRI,K(t, ξ), (3.4) Chapter 3. Modeling 14 Figure 3.1: Fifty-second electrocardiography (ECG) waveforms, time tracings of R-R interval (RRI), ECG-derived respiration (EDR), and the corresponding spectra in lighter level of anesthesia (A, B, C) and deeper level of anesthesia (D, E, F). Whether in lighter (A) or deeper level of anesthesia (D), it is hard to read the IHR information from raw ECG waveform by naked eyes. Time tracings (B, E) of RRI and EDR and their power spectra (C, F) show that RRI is more non-rhythmic in lighter level of anesthesia then in deeper level. The frequency bands around the instantaneous frequency f r and its multiple frequencies (the unshaded area) constitute the rhythmic component. The rest (shaded area) constitutes the non-rhythmic component. dB = decibel Chapter 3. Modeling 15 which is implemented by the curve extraction algorithm considered in [49]. Next, we will calculate the rhythmic component power, denoted as Pr(t), as the sum of the power inside the bands around fr(t) on the tvPS of RRI(t). In particular, (cid:90) nfr(t)+0.01Hz nfr(t)−0.01Hz Pr(t) := MRRI,K(t, ξ)dξ. Here the width of the band is chosen to be 0.01Hz in an ad hoc way. We mention that when the signal is of broadband spectrum, there is no dominant curve that can be visualized, as shown in Fig.3.1. In this case, the most dominant curve defined in (3.4) from the time-frequency representation is viewed as the rhythmic component. The non-rhythmic component power is defined as the power within the high frequency subtracted by the rhythmic component power. Lastly, NRR is defined intuitively as the ratio of the non-rhythmic component power to the rhythmic component power: (cid:18) tvHF(t) − Pr(t) (cid:19) . Pr(t) NRR(t) = log10 Through defining power within absolute boundary as Pr(t), NRR can differentiate the level between "rhythmic" and "non-rhythmic". In this thesis, the variables affixing (t) mean that they are functions of time. It is important to notice that these sorts of time-varying parameters describe the momentary behavior of the signal. Since this instantaneous in- formation cannot be captured by the classical power spectral analysis, we would expect to gain more insight into the IHR signal by considering our definitions. 3.5 Non-rhythmic Component vs. Stochastic Process For the purpose of this thesis, the term non-rhythmic is defined as any signal that appears "irregular", "noisy", or "as random". This poses two interesting questions. First, does this mean that we should treat all non-rhythmic signals as a stochastic process? Second, will we be able to model the non-rhythmic signal as an auto-regressive moving-average (ARMA) model or other stochastic process in order to analyze it using parametric meth- ods? However, I would like to also remind that we should consider these issues using our medical knowledge. One of my hypotheses in the present study is that the non-rhythmic signal reflects highly structured and complex activity of the human brain. Although it is too early to determine that the human brain is a "deterministic" or "stochastic" system, when awake, the human brain can retain memory for longer periods of time to execute complex tasks. The thought or idea in the brain lasts longer too. Therefore, the brain exhibits non-rhythmic activity Chapter 3. Modeling 16 when its inner state is longer lasting. However, the above statements would contradict the stochastic process behavior that presents no memory effect in a Gaussian random process. That is, the autocorrelation function is a Dirac's delta function: (cid:88) l∈Z Ryy(l) = y(n)¯y(n − l), E[RRyy(0)] = σ and E[RRyy(k)] = 0 for k > 0 when y is a Gaussian process. Although non-rhythmic signals may appear similar to a stochastic process, the human brain that shows non-rhythmic activities should result in a more structured state. There- fore, a non-rhythmic signal in its natural state is not completely "stochastic". This is why I would like to use the term "non-rhythmic". My second hypothesis is that there are two components, non-rhythmic and rhythmic, that are contained in the "rhythmic-to-non-rhythmic" phenomenon. It is unknown how these two components coexist but one simple way to model their co-existence would be superposition. Regardless if we fully understand how the two components coexist, the fact that stochastic process is fundamentally a "white-noise-driven" process means it doesn't match the "rhythmic component plus non-rhythmic component" concept. The above technical issues regarding parametric modeling for "rhythmic-to-non-rhythmic" phenomenon needs to be carefully addressed in the future work. To narrow down the problems I have to solve in my thesis, I decided to choose a non-parametric method as a starting point. Still, statistical signal processing technique is a viable direction for my future studies. 3.6 Summary In summary, the proposed adaptive harmonic model (3.2) describes a fundamental physi- ologic property: the rhythmic component possesses frequency modulation and amplitude modulation. However, the law of physiology and chemistry sets a limit to these modula- tions. From the previously mentioned theory to the modeling and quantification of the NRR index, I keep seeking the time-varying ability to capture the momentary dynamics, which is essential to the dynamic nature of clinical anesthesia. This feature is also fundamentally distinct to theories of other signal analysis methods. After establishing the theoretic part, I will focus on the next problem: the time-varying spectrum. Chapter 4 Time-Frequency Analysis, Reassignment, and Synchrosqueezing 4.1 Time-Frequency Analysis Chapter 3 provided strong evidence that in order to execute the quantification of NRR index within anesthesia the adaptive harmonic model requires a strong time-varying spec- trum. A recently developed time-frequency analysis technique is called the multitaper Synchrosqueezing transform. I believe that this technique is the most useful for my pro- posed study. Time-frequency analysis is a type of signal processing tool that generalizes the power spectrum to provide time-varying ability. This type of analysis generates a series of power spectra that has the ability to represent instantaneous spectra. Short-time Fourier trans- form (STFT) and continuous wavelet transform (CWT) are two types of well known tools for biomedical signal analysis[43, 50 -- 53]. Both of these tools are able to capture the instan- taneous frequency information of the biological signal. Huang and Chan were the first to employ time-frequency analysis to analyze the dynamic changes in IHR during anesthesia [46]. Both STFT and CWT require window functions to "focus" or "localize" the signal. Take STFT as example: (cid:90) ∞ −∞ V (h) X (t, ξ) = X(s)h(t − s)e−i2πξsds (4.1) 17 Chapter 4. Synchrosqueezing Transform 18 where h is a window function chosen by the user, for example, the Gaussian function h(t) = e−t2. Note that X(τ )h(t − τ ) is the piece of signal around t chosen by the window function. We then define the spectrogram of X(t) [50] as X (t, ξ) =(cid:12)(cid:12)V (h) X (t, ξ)(cid:12)(cid:12)2. G(h) (4.2) From equation 4.1, it is apparent that choosing different window functions causes different results. The influence of the chosen window function in relation to the STFT is no less than the signal per se. In particular, the window function imposes a resolution limit which relates to the well-known Heisenberg's uncertainty principal. The larger the time window h is, the higher the spectrum resolution is, but with low temporal resolution, and vise versa. The window function also influences other TF methods, such as CWT, and Smoothed Pseudo Wigner-Ville Distribution, to name a few. Rather than fiddling with various TF methods, various window functions, and different sizes, it is highly advised to choose a technique that alleviates the above issues. 4.2 Reassignment To cope with the resolution limitation of time-frequency analysis, it is possible to further use the phase information ϕ(t, ξ), which is discarded in the classical STFT spectrogram (4.2) and scalogram[50, 51]. Kodera et al. first proposed this technique[54] by calculating the instantaneous frequency ∂tϕ(t, ξ) and group delay ∂ξϕ(t, ξ). Auger and Flandrin coined this technique as the reassignment and further generalized it to "Fourier transform based" time-frequency analysis[50]. The reassignment technique provides a sharper representa- tion, which improves the readability of the STFT spectrogram. Reassignment is helpful in visualizing the dynamics of the biological signal that contains multiple nonstationary components in the TF plane (Fig. 4.1). The phase information hidden in STFT contains information that explains why time- frequency domains are spreaded out.[55]. To further eliminate this issue, we can shift the STFT coefficients accordingly to the reassignment rules determined by the phase informa- tion. The temporal reassignment rule and frequency reassignment rule are as follows: Chapter 4. Synchrosqueezing Transform 19 Figure 4.1: A side-by-side comparison between STFT spectrogram (A) and Reassign- ment spectrogram (B). The tracing of original R-R interval signal recorded in anesthesia is denoted as red line. Chapter 4. Synchrosqueezing Transform 20 (cid:40) (cid:98)tt,ξ := t + (cid:60) (cid:40) (cid:98)ξt,ξ := ξ − (cid:61) (cid:41) (cid:41) , V (T h) X (t, ξ) V (h) X (t, ξ) V (Dh) (t, ξ) X V (h) X (t, ξ) where (cid:60) denotes the real part and (cid:61) denotes the imaginary part and h, T h, and Dh are the three related window functions. Here T h is the time-ramped window used to determine the temporal reassignment rule and Dh is the differential window used to de- termine the frequency reassignment rule. The reassignment is then achieved by applying the reassignment rules: RS(h) X (t, ξ) := (cid:90) ∞ (cid:90) ∞ −∞ −∞ X (s, η)δ(s −(cid:98)ts,ξ)δ(η −(cid:98)ξs,ξ)dsdη, V (h) where we move the spectrogram coefficient V (h) location(cid:98)ts,ξ and (cid:98)ξs,ξ according to the reassignment rules. X (s, η) at time s and frequency η to a new The reassigned spectrogram is a valuable tool for signals that contain dynamic information of anesthesia in my current study[43]. However, aside from the visual information repre- sented in the TF plane, it is also desirable to execute further analysis of the signals, such as isolating a component in TF plane, or performing reconstruction in time domain. Syn- chrosqueezing transform is a recently developed reassignment technique that can provide this possibility. 4.3 Synchrosqueezing Transform Synchrosqueezing transform is a special kind of reassignment technique[44, 45, 56]. This technique uses a light-weight mathematical model (4.5). The light-weight mathematical model helps with further processing of signals, such as isolating a component in the TF plane or reconstruction in time domain. For my present study, Synchrosqueezing transform can be used to present non-parametric generalizations of the power spectrum. The Synchrosqueezing transform is able to provide the time-varying power spectrum (tvPS) which reveals the instantaneous changes that happen in IHR. In addition, comparing with traditional STFT and wavelet, it is visually more informative, more resistant to several kind of noise, and more resistant to the influ- ence of chosen window function [44, 45, 49]. Mathematical works have supported these abilities, as well as the ability of inverse transform. Luckily we do not need to impose too Chapter 4. Synchrosqueezing Transform 21 Figure 4.2: A side-by-side comparison between STFT spectrogram (A) and Reassign- ment spectrogram (B). The tracing of original R-R interval signal recorded in anesthesia is superimposed as red line in panel B much effort on choosing parameters as using the Synchrosqueezing transform technique can easily capture the oscillatory activity. The detailed information on the algorithm leading to tvPS is as follows. It has been known that the phase information hidden in STFT contains more information that explains why the STFT spectrogram is spreaded out. A solution that could help remedy this issue is by shifting the STFT coefficients according to reassignment rules determined by the phase information. The frequency reassignment rule is as follows: Chapter 4. Synchrosqueezing Transform Ω(t, ξ) := −i∂tV (h) 2πV (h) X (t, ξ) X (t, ξ) 22 (4.3) where ∂t means the partial derivative with related to t. The Synchrosqueezing transform is then achieved by applying the reassignment rule: (cid:90) ∞ (cid:90) ∞ X (t, η)δ(ξ − Ω(t, η))dη, V (h) S(h) X (t, ξ) := −∞ −∞ where for each fixed time t, we move the STFT coefficient V (h) new location according to the reassignment rules Ω(t, ξ). X (t, η) at frequency η to a One way that the Synchrosqueezing transform is unique to the reassignment technique is that the Synchrosqueezing transform has the ability to reconstruct back to the time domain. Suppose the signal X(t) analyzed by Synchrosqueezing transform is: where T (t) > 0, A(t) > 0 and φ(cid:48)(t) > 0. X(t) = T (t) + A(t) cos(2πφ(t)), (4.4) In other words, X(t) comprises of an oscillatory component and a trend component. In addition, the amplitude modulation and frequency modulation of the oscillatory com- ponent is under the condition as follows: 0 < A(t) < T (t), φ(cid:48)(t) > 0, A(cid:48)(t) ≤ φ(cid:48)(t), φ(cid:48)(cid:48)(t) ≤ φ(cid:48)(t), T (t)(cid:48) ≤  for all t. (4.5) (4.6) Relevant results of the Synchrosqueezing transform can be found in literatures[49, 57 -- 60]. A recent review article discussed in depth on the relation of the TF analysis, the reassignment and Synchrosqueezing technique[61]. 4.4 Multitaper Estimation From a non-parametric spectral estimation perspective, Thompson proposed a powerful multi-window technique that improves spectral estimation of the random process[62, 63]. Thompson uses the discrete prolate Spheroidal sequence to define a series of window func- tion, which are orthogonal to each other, that have the same length in time domain. The average of the spectra from multiple windows reduces the spectrum variance but also prevents the bias of spectral estimation induced by varying window size. One strong Chapter 4. Synchrosqueezing Transform 23 advantage of Thompson's spectrum is that we are able to estimate the spectrum of a stationary random process without resort to parametric estimations. The principle of Thompson's technique has been extended to the field of nonstationary signal. Multitaper in time-frequency analysis is designed to be used with Hermite functions as the sequence of multiple windows that are orthogonal to each other in time-frequency plane. The final spectrogram estimation can then be obtained from the average of multiple realizations. Taking the STFT spectrogram as an example, the multitaper spectrogram can be defined as follows: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 K K(cid:88) k=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 V (hk) X (t, ξ) . QX,K(t, ξ) := The multitaper spectrogram and multitaper scalogram were shown to make success for nonstationary stochastic signal[64]. To achieve both high resolution for deterministic non- stationary signal and low variance for stochastic nonstationary signal, Xiao and Flandrin combined the reassignment technique and multitaper technique[47]. From the background information mentioned in Chapter 2 and the model in Chapter 3, it can be seen that the IHR under anesthesia should contains both non-stationary component and stochastic component. We have reported the advantage of using multitaper time- frequency reassignment in extracting these dynamic features[43]. 4.5 Multitaper Synchrosqueezing Spectrogram Synchrosqueezing transform responds well to signals from different types of stochastic process. One dilemma we face is understanding the true mechanism in biological signal. Since the quantification of the non-rhythmic component is an important part of our current model(3.4), we have decided to apply the multitaper method to further optimize the performance of the Synchrosqueezing transform X(t). We have chosen the multitaper Synchrosqueezing spectrogram to achieve the tvPS as follows: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 K K(cid:88) k=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 MX,K(t, ξ) := S(hk) X (t, ξ) , where K > 0 is the number of windows chosen. We choose the the k-th Hermite function as hk. For my current study, I will use K = 10 and the first ten Hermite window functions will be hk, k = 1, . . . , 10 when deriving tvPS. Chapter 4. Synchrosqueezing Transform 24 Figure 4.3: The time-varying power spectrum of instantaneous heart rate (IHR) as data used in figure 4.2, realized by multitaper Synchrosqueezing spectrogram, showing rich dynamic features of human body in anesthesia. As shown in figure 4.2, the readability is better than STFT spectrogram, and also better than the reassigned spectrogram. The advantage of using a multitaper Synchrosqueezing spectrogram is summarized as follows: 1. Robust to several different kinds of noise, which might be slightly nonstationary. 2. Visually informative. When the signal is rhythmic, that is, satisfying the model (3.1), a dominant curve following the IF can be seen on the multitaper Synchrosqueezing spectrogram; otherwise the multitaper Synchrosqueezing spectrogram is blurred. 3. Adaptive to the data so that its dependence on the chosen window function is weak. We compared various methods of TF analysis in Table4.1 and in simulated data analy- sis(Fig.4.4). 4.6 Adaptive Filter and Parametric Estimation Another kind of techniques based on parametric modeling for time series data analysis is the adaptive filter, including the famous Kalman filter, and the parametric spectral estimation[65]. The adaptive filter is designed to use a cost function to represent the weighted mean square error. The goal is finding a set of parameters ai to minimize the cost function. ai = argmin E{e2[n]} Chapter 4. Synchrosqueezing Transform 25 Figure 4.4: An artificial data, as a summation of two time-varying components and a Gaussian white noise, and its TF analysis realized by STFT spectrogram, Syn- chrosqueezed spectrogram and the multitaper Synchrosqueezed spectrogram. From 125 s to 250 s, it is apparent that Synchrosqueezing technique improves the readability of the two components, and multitaper technique refines further. If we have adequate knowledge about the signal, we can obtain the optimal point on the performance surface of the cost function by using partial differential. E{e2[n]} = 0 ∂ ∂ai It is apparent that TF reassignment is fundamentally different from adaptive filter. TF reassignment is a non-parametric deterministic signal processing technique, whereas the adaptive filter is a parametric technique for stochastic signal processing. Regardless of the difference, both techniques use the same mathematical tools to achieve their goals. Kalman filter in essence is a real-time prediction-correction algorithm that provides para- metric estimation in the sense of weighted least-square fitting[62, 65]. As the physiologic mechanism underlying rhythmic-to-non-rhythmic phenomenon is not adequately under- stood, a technique which requires both elaborate model construction and careful parame- ter selection is not my first-line choice for my research problem. Besides, From Chapter 3, it shows that the NRR index has a high relation to oscillation. Therefore, for my study, Chapter 4. Synchrosqueezing Transform 26 Table 4.1: Summary of abilities of TF methods. Details can be found in references cited in this chapter. X: unable; V: feasible; ?: questionable Reassignment SST STFT CWT Instantaneous frequency estimation good good fair fair Visual readability good good fair fair Component reconstruction Robust to noise X ? V V V good fair fair the output of the Kalman filter as a matrix of state variables may not be as intuitive as the spectrum technique for data exploration and interpretation at this stage. Parametric spectral estimation has a complete theory of modeling based on stochastic process[62]. One critical issue of the parametric estimation is that the mechanism of the signal we are facing in the present study is unknown, and is constantly changing. The para- metric spectral estimation has a lower tolerance to incorrect modeling. In other words, it is crucial to ensure that the correct model are constructed to generate reasonable parametric spectral estimation. In addition, parametric spectral estimation does not behave well when the signal is a superposition of multiple components. The rigidity of model construction contrasts well with the flexibility of Synchrosqueezing: flexible in window selection and adaptive to signal. Certainly, it is possible that the technique realizing time-varying power spectrum in the future may belong to the family of parametric spectrum. 4.7 Common Condition in Physiology and Mathematics The most unique characteristic of the Synchrosqueezing transform that is relevant to this project is the mathematical condition 3.2 which shows that the frequecy modulation and amplitude modulation are limited. The condition used on the Synchrosqueezing transform is exactly the same condition as the adaptive harmonic model in order to describe this rhythmic physiological behavior. Therefore, the Synchrosqueezing transform matches per- fectly with the adaptive harmonic model for us to quantify the "rhythmic-to-non-rhythmic" phenomenon that was previously discussed. Chapter 5 Performance Evaluation using Clinical Data 5.1 Clinical Database In my hopes to investigate the clinical value of NRR index, I have collected and analyzed a clinical database. The prospective observational study spanned the whole period of typical surgical anesthesia. ECG, BIS index, and the end-tidal concentration of sevoflurane were all recorded continuously. Anesthetic events, including loss of consciousness (LOC), skin incision, first reaction of motor movement during emergence period, and return of consciousness (ROC), were registered with precise time stamps. After receiving approval from the Institutional Review Board (Taipei Veterans General Hospital, Taipei, Taiwan), we enrolled 31 female patients. All patients were American Society of Anesthesiologists (ASA) physical status of I or II and were scheduled for la- paroscopic gynecological surgery undergoing general anesthesia. We obtained informed consent in paper from each of our patients. All ASA I or II status patients are generally under healthy conditions. It is important to investigate the changes of the NRR index under the normal physiological conditions. We are aware that some diseases can affect HRV (and IHR) untowardly. Thus, patients that are excluded were those younger than 20 years old or older than 55 years old, have a body mass index below 19 or higher than 30 and have either cardiovascular disease, arrhythmia, intake of medication that affects the neurological and cardiovascular system, potential cancer, diabetes mellitus, or anticipated difficult airway. All these surgeries were performed between the hours of 8 a.m. and 7 p.m. 27 Chapter 5. Performance Evaluation 28 5.2 Anesthesia Protocol The standard anesthetic monitoring (HP agilent patient monitoring system) was applied to all patients. This included ECG, pulse oximeter (SpO2) and non-invasive blood pressure. A BIS electrode (BIS-XP sensor) and the ECG recorder (MyECG E3-80; Micro-Star Int'l Co., New Taipei City, Taiwan) were applied to the patients prior to anesthetic induction. After the initial step of cleaning the skin with rubbing alcohol, the BIS sensor is applied to the right side of the forehead. As a standard practice, all anesthetic managements, including the administration of medications, and airway management were under the discretion of the anesthesiologist. After intravenous infusion of lactated Ringer's solution begins, we started anesthetic induction with fentanyl 2.5 -- 3 µg/kg. Next, right after preoxygenation, hypnosis was induced with propofol 2 -- 2.5 mg/kg. LOC was then assessed by no response to verbal command. Cisatracurium 0.15mg/kg was used to facilitate tracheal intubation. Subsequently, mechanical ventilation was started from volume control mode with oxygen- gas-sevoflurane mixture as low flow anesthesia. The respiratory rate and tidal volume of the ventilator was adjusted to maintain an end-tidal carbon oxide (ETCO2) that ranges between 35 and 40 mmHg, and to keep the peak airway pressure lower than 25 mm Hg. The laparoscopic skin incision was made by a surgical blade. The adequacy of anesthetic depth (during the whole period of surgery) was determined by the BIS index (< 60) and based on the anesthesiologist's judgment. If the anesthesiologist determined an inadequate level of analgesia, a bolus dose of fentanyl would be given to the patient. Toward the end of the surgery, the peritoneal gas was deflated and the patient's body position recovered from Trendelenburg's position to level-supine position. As the wound closure began, the anesthetic gas was on a consistent decrease. Muscle relaxation was reversed with a com- bination of neostigmine 0.05 mg/kg and glycopyrrolate 0.01 mg/kg during controlled ven- tilation. Once the patient regained spontaneous breathing the controlled ventilation was halted. Patients that exhibited inadequate spontaneous breathing (ETCO2 >50 mmHg or SpO2 <95%) were assisted with manual positive ventilation, otherwise the patients continued to breath spontaneously until regaining consciousness without the interference of positive-pressure ventilation.ECG and BIS are the main electrophysiological signals we analyze for this project. When we observed the emergence period, physical contact to the patient was avoided and minimized if necessary to prevent the interference on the sensors of ECG and BIS monitor. Any unnecessary interference to patient's respiration from the manual (bag) ventilation was avoided also. When the patient showed adequate consistency in their spontaneous breathing, we removed the endotracheal tube. If in any case there was inadequate ventilation or upper airway obstruction, corrective actions were taken that include either mask ventilation or nasopharyngeal airway insertion. During the emergence period, I carefully assessed first reaction as any first visible motor reaction such Chapter 5. Performance Evaluation 29 as movement of the arms or legs, coughing, or grimace. ROC was assessed through the patient's ability to open their eyes and follow simple procedural commands. 5.3 Data Acquisition The BIS index was continuously recorded from an Aspect A-2000 BIS monitor (version XP, Host Rev:3.21, smoothing window: 15 seconds; Aspect Medical Systems, Nattick, CA, USA) connected to a laptop computer (Asus Corp., Taipei, Taiwan). During surgery, corrective measures were taken to improve signal quality of the BIS sensor when the signal quality index was lower than 50; otherwise the BIS data during this period was discarded. The raw limb lead ECG sampled at 1000Hz and 12-bit resolution, was recorded for off-line analysis. The clocks on the laptop and ECG recorder were synchronized with a time accu- racy of ±1 second. Time stamps of BIS record were provided by the laptop. The inhaled and end-tidal concentrations of anesthetic gas, which was sampled from the connection piece close to the endo-tracheal tube and detected by the gas analyzer on a Datex-Ohmeda S/5 anesthesia machine (GE Health Care, Helsinki, Finland), were also recorded on the laptop. All data was recorded without any interruption until the patient regained ROC. All data collected for analysis of sevoflurane concentration correlation and anesthetic events, including skin incision, first reaction and ROC, are under single ventilation mode. There were no transitions from the mechanical ventilation to the spontaneous ventilation or vice versa in the data. However, during LOC, the transition from the spontaneous ventilation to mechanical ventilation is possible. The offline ECG signal was analyzed in several steps. The R peak detection was auto- matically determined by taking lead I, II and III ECG signals into account to improve accuracy. The ectopic beats were removed and interpolated from the data to eliminate incorrect R peaks and ectopic beats through visual verification. When electrocauteriza- tion is severely interfered by the ECG, the data segment were also discarded. Whenever electro-cauterization severely interfered the ECG, this data segment was discarded. IHR was then derived from the R peaks through the cubic spline interpolation. RRI were resampled to be equally spaced at 4 Hz from IHR for subsequent analyses (Fig.2.1). In order to correct the hysteresis between the end-tidal sevoflurane concentration and the anesthetic effect on the brain, the estimated effect-site sevoflurane concentration (Ceff) was derived from the end-tidal sevoflurane concentration (Cet) by the following first order differential equation. This equation is based on the pharmacokinetic-pharmacodynamic modeling[66]: Chapter 5. Performance Evaluation 30 dCeff dt = Ke0(Cet − Ceff) where the constant Ke0 was assumed to be constant for all patients and defined as 0.20 /min according to previous studies [8]. 5.4 Statistical Analysis We treated BIS and ECG-derived continuous indices, including NRR, tvHF, tvLF, tvLHR and HR, as anesthetic depth indices. The performances of the indices were evaluated in two ways: the ability for these indices to predict anesthetic events and their correlations with effect-site sevoflurane concentration during emergence period. Because the concentration of anesthetic gas consistently decreased and the influence from surgery was relative minor, we considered the emergence period for correlation analysis. A p value less than 0.05 would be considered as statistically significant. Multiple significant tests were calculated by using Bonferroni correction. Statistical results are expressed as mean (standard deviation [SD]). Prediction probability (PK) analysis is a versatile statistical method for measuring the performance of an anesthetic depth index [67]. A value of one means that the observation is always correctly predicted, a value of 0.5 indicates that the observation is predicted at a 50/50 chance. The ability to predict anesthetic events can be investigated by using the serial prediction probability analysis[67]. The correlation with sevoflurane concentration employed PK analysis and Spearman rank correlation. Estimation of PK and its standard error was obtained through the Jackknife method. Prior to using the null hypothesis, any PK value less than 0.5 would be changed into one minus the PK value[67]. 5.5 Serial Prediction Probability Analysis in Predicting Anes- thetic Events To evaluate the performance of the anesthetic indices in predicting the anesthetic events, serial PK analysis was designed as follows. The following timestamps are used as the baseline: one minute before LOC, five seconds before skin incision, three minutes before the first reaction and three minutes before ROC. The serial PK analysis is performed index on baseline timestamp versus indices on as successive PK analyzes of data pairs: subsequent successive timestamps spaced by five seconds. As a result, the out of the serial PK analysis is a sequence of time-varying PK values which reveals the temporal relation between anesthetic events and the indices. By plotting the serial PK value and its standard error bar, it becomes possible for us to do hypothesis test simply with the naked eye. If Chapter 5. Performance Evaluation 31 there is significant difference (p < 0.05), it can be established if 1.5 times of their standard error bars do not overlap [67, 68]. 5.6 Correlations with Sevoflurane Concentration Correlation between the above indices and sevoflurane concentration was investigated in the emergence period. We chose the time interval where the sevoflurane concentration is monotonically and continuously decreased. The period of spontaneous breathing is defined from the start of adequate spontaneous breath to ROC. We employed the PK analysis and Spearman rank correlation to analyze performance of indices sampled every 4 seconds. The results were tabulated as weighted averages according to each patient's data length. The bootstrap method was used to calculate the 95% confidence interval of Spearman correlation based on 10,000 samplings. A value of Spearman correlation closer to 1 is better since we expect that the indices will increase to correlate with the decrease of sevoflurane concentration. 5.7 Algorithm of Serial Prediction Probability (sPK) Anal- ysis In this section we detail a variation of the Prediction probability (PK) analysis [67], referred to as Serial Prediction Probability (sPK) analysis. sPK is employed in this study to evaluate the prediction performance of the BIS and ECG-derived continuous indices, including NRR(t), tvHF(t), tvLF(t), tvLHR(t) and HR. We consider K different measurement times, which are the times when we record the anesthetic indicators. Then we determine two special timestamps, one is referred to as the event time, indicating the happening of the event, for example, the first reaction, and one is referred to as the base time, which is T > 0 seconds before the event time. Here T depends on the event we are studying. Collect N subjects and record the measurement times and the event time of each individual. Note that the measurement times and the event time and are different among individuals, and we denote the k-th measurement time of the n-th subject as tk,n, k = 1, . . . , K, and the base time as t0,n. Denoted as xk n the anesthetic indicator (e.g. BIS) of the n-th subject recorded at time tk,n, and denoted as yk n the associated outcome (e.g. first reaction) at time tk,n. To evaluate the prediction power of an anesthetic indicator at the k-th measurement time, we take the following two datasets. At the measurement time tk,n, we record the sam- ple anesthetic indicators {xk N}, where each N} and the sample outcomes {yk 1, . . . , xk 1 , . . . , yk Chapter 5. Performance Evaluation 32 n, yk pair (xk n) is recorded from the n-th subject. Denote Zk := {(xk N )} as the measurement data at the k-th measurement time. Moreover, at the base time t0,n, we record the sample anesthetic indicators {x0 N} and the sample outcomes {y0 1), . . . , (x0 as the data at the base time. N}, where each pair is recorded at the base time. Denote Z0 := {(x0 1, yk 1 ), . . . , (xk 1, . . . , x0 1, . . . , y0 N , yk 1, y0 N )} N , y0 For a given individual, there are five possible relationships between pairs of anesthetic indicator and outcome at the k-th measurement time and those at the base time: 1. (xk n, yk n) and (x0 m, y0 m) are concordant if they are rank ordered in the same direction; n) and (x0 m, y0 m) are disconcordant if they are rank ordered in the reverse di- 2. (xk n, yk rection. 3. (xk n, yk n) and (x0 m, y0 m) tie in x if xk n = x0 m while yk 4. (xk n, yk n) and (x0 m, y0 m) tie in y if yk n = y0 m while xk n (cid:54)= y0 m; n (cid:54)= x0 m; 5. (xk n, yk n) and (x0 m, y0 m) tie in both x and y if xk n = x0 m and yk n = y0 m. Usually the anesthetic indicator is of finer scale and the outcome is of coarser scale. Thus, unlike the usual notion of correlation, in the PK analysis the tie in x is undesirable but we tolerate the tie in y and the tie in both x and y. Based on this notion, we define the serial PK index in the following way. Denote Pc(k), Pd(k) and Ptx(k) the respective probabilities that two pairs of (x, y) independently drawn from Zk and Z0 with replacement are concordant, disconcordant and tie in x at the k-th measurement time. The serial PK index at the k-th measurement time is denoted as PK(k) = Pc(k) + 1 2 Ptx(k) Pc(k) + Pd(k) + Ptx(k) . The PK(k) value is interpreted in the same way as that in the PK analysis. A value of one means that the indicator always correctly predicts the observed depth of anesthesia, a value of 0.5 means that the indicator predicts no better than 50/50 chance, and a PK(k) value less than 0.5 means that the indicator predicts inversely. To be more precise, a zero value also means a perfect prediction but in a reversed direction. The estimation of the serial PK and its standard error was obtained by the Jackknife method. Before the null hypothesis, the serial PK value less than 0.5 was converted into one minus the serial PK value. Chapter 6 Results from Clinical Database 6.1 Visual Information in TF Plane Abundant information can be visually read from the tvPS of IHR. Various features were determined to reflect the physiologic dynamics of human body responding to various anes- thetic events (Fig. 6.1). It is demonstrated (Fig. 6.2,6.3) that the tvPS is more concentrated on deeper anesthetic levels and more scattered on lighter anesthetic levels. This then created the hypothesis on trying to understand the "rhythmic-to-non-rhythmic" phenomenon. This phenomenon is consistently seen in most study cases. During spontaneous breathing (Fig.6.3), the dominant curves can be seen in the first portion (usually on the left hand side) of the tvPS graph. Meanwhile, less dominant curves can be identified in the second portion (usually the right hand side). The dominant curve represents the typical rhythmicity feature. It is also important to observe that the tvPS becomes more non-rhythmic prior to the appearance of first reaction. (Fig. 6.3) Painful surgical stimulations during anesthesia are referred to as noxious stimulation be- cause an unconscious patient under anesthesia cannot report pain sensation. Two of the most intense noxious stimulationthat occur in anesthesia and surgery are endotracheal in- tubation and skin incision. Clinical anesthesiologist uses the increase of heart rate and blood pressure as indicators for the level of noxious stimulation. In my study, we have found the visual information of tvPS that reveals the influence of noxious information where the low frequency (LF) power increases immediately after noxious stimulation, and wanes soon. The "LF surge phenomenon" appeared in all study cases. Although the main goal of my study is on investigating NRR index, it can be seen that the potential value of LF power as an "noxious index" deserves to be investigated too. 33 Chapter 6. Results 34 Figure 6.1: The time-varying spectrum of R-R peak interval (RRI) recorded during anesthesia showing rich dynamic features of human body in anesthesia. The RRI tracing is superimposed as red line These visual findings are quantified by NRR and HRV indices, such as tvHF, tvLF and tvLHR for subsequent analyses. 6.2 Multiple Component Phenomenon Occasionally, multiple independent oscillatory components hidden in IHR can be seen from the tvPS. To further expand on this statement, the term independent is defined as the frequency of one component of which the modulation is unrelated to other com- ponents. Therefore, this is definitely not the harmonics (Fig.6.8, 6.9, 6.10, 6.11). The "multiple component phenomenon" are an interesting finding, which demonstrates the dynamic physiology of human body undergoing anesthesia. Apparently, it is nearly impossible to identify the appearance of multiple component from raw IHR waveform using naked eye. The tvPS is good at presenting "multiple compo- nent phenomenon", but it is a difficult task for classical power spectrum. The "multiple component phenomenon" appeared in one thirds of the study cases. The moment that it appeared and disappeared seems random and unpredictable. Even so, this phenomenon Chapter 6. Results 35 Figure 6.2: A representative data from a patient under anesthesia and controlled venti- lation. The 1200 seconds R-R interval (RRI) is superimposed on its time-varying power spectra (A) as a blue solid line. The corresponding electrocardiography derived respira- tion (EDR) is superimposed on its time-varying power spectra (B) as a green line. Panel C shows simultaneous recorded Bispectral index (BIS), effect-site sevoflurane concentra- tion (Sevo) and non-rhythmic to rhythmic ratio (NRR). Arrow 1: inadequate level of anesthesia was noted and the sevoflurane concentration was raised. Sevoflurane concen- tration reached top level after 300-second. Arrow 2: both BIS index and NRR reached minimum. There is a dominant straight line around 0.167 Hz in the panel A, whose fundamental frequency is the same as the rate of the ventilator, representing the en- trainment of RSA by the controlled ventilation. The time-varying power spectra of RRI showed rhythmic-to-non-rhythmic transition, corresponding to varying anesthetic depth, whereas the time-varying power spectra of EDR was almost unchanged under controlled ventilation. Chapter 6. Results 36 Figure 6.3: A representative data obtained from the same patient under spontaneous breathing in emergence period. The 1200-second R-R interval (RRI) is superimposed on its time-varying power spectra (A) as a green solid line. The corresponding electrocar- diography derived respiration (EDR) is superimposed on its time-varying power spectra (B) as a blue solid line. Panel C shows simultaneous recorded Bispectral index (BIS), effect-site sevoflurane concentration (Sevo) and non-rhythmic to rhythmic ratio (NRR). Arrow 1: endotracheal suction and extubation, Arrow 2: first reaction, Arrow 3: regain of consciousness. Under spontaneous breath, RRI and EDR are similar in high frequency region of spectrograms. The rhythmic component and its harmonics are time-varying in both time-varying power spectra of RRI and EDR. With the overall decrease of sevoflu- rane, both RRI and EDR exhibited a rhythmic to non-rhythmic transition. The RRI and EDR became non-rhythmic before the first reaction.(From ref.[69]) Chapter 6. Results 37 Figure 6.4: A representative data from a patient under anesthesia receiving skin incision for her surgery. The 750-second R-R interval (RRI) is superimposed on its time-varying power spectra (A) as a green solid line. The corresponding electrocardiography derived respiration (EDR) is superimposed on its time-varying power spectra (B) as a blue solid line. Panel C shows simultaneous recorded Bispectral index (BIS), effect-site sevoflurane concentration (Sevo) and non-rhythmic to rhythmic ratio (NRR). is related to specific anesthetic events like only appearing when the patient is undergoing anesthesia. Although mechanism underlying this phenomenon is beyond the goal of my present study, it still presents an interesting research topic for the future. 6.3 Ability to Predict Anesthetic Events In the serial PK analysis for the first reaction (Fig. 6.14), the NRR is ahead of BIS (p < 0.05 for 20 seconds). The NRR was able to predict the first reaction 30 seconds Chapter 6. Results 38 Figure 6.5: tvPS of IHR during emergence period from the last three consecutive cases (no.1) showing rhythmic-to-non-rhythmic phenomenon. BIS index (golden yellow tracing) and IHR (blue tracing) are superimposed. Figure 6.6: tvPS of IHR during emergence period from the last three consecutive cases (no.2) showing rhythmic-to-non-rhythmic phenomenon. BIS index (golden yellow tracing) and IHR (blue tracing) are superimposed. Chapter 6. Results 39 Figure 6.7: tvPS of IHR during emergence period from the last three consecutive cases (no.3) showing rhythmic-to-non-rhythmic phenomenon. BIS index (golden yellow tracing) and IHR (blue tracing) are superimposed. Figure 6.8: The time-varying power spectrum of instantaneous heart rate (IHR) recorded during anesthesia showing three independent oscillatory components, revealing rich dynamic features of human body in anesthesia. Chapter 6. Results 40 Figure 6.9: The IHR (blue tracing) was recorded during spontaneous breathing. The second component appeared shortly (200 -- 300 s). Figure 6.10: The IHR during spontaneous breathing showed "rhythmic-to-non- rhythmic" changes with BIS index (golden yellow tracing), accompanied with multiple components during "rhythmic" period. Chapter 6. Results 41 Figure 6.11: Another IHR data (blue tracing) during spontaneous breathing showed "rhythmic-to-non-rhythmic" changes with BIS index (golden yellow tracing), accompanied with multiple components during "rhythmic" period. in advance (PK > 0.90). At the instance of first reaction (0 seconds), both the NRR and BIS (both PK maxima > 0.95) were significantly better than other parameters. The time-varying HRV indices and HR (PK < 0.83) provides significantly worse results than the NRR. In the serial PK analysis for skin incision (Fig. 6.4), tvLF reaches maximum (PK > 0.95) roughly at 30 seconds after skin incision. It reflects skin incision best, followed by tvLHR (PK < 0.85). The tvLF is significantly better than BIS (PK < 0.55) and NRR (PK < 0.65). In the serial PK analysis for LOC (Fig. 6.12), BIS reflects perfectly (PK = 1) 50 seconds after LOC. LOC is also associated with a decrease in tvLF, tvHF and HR. Nowever, NRR does not reflects LOC well. In the serial PK analysis for ROC (Fig. 6.15), BIS is the best index, and surpasses NRR, HR and the time-varying HRV indices significantly. Due to the inadequate data quality, 1 patient has been eliminated from our data in regards to the serial Pk analysis of the initial reaction. For our study, when we analyzed the serial Pk of skin incision, LOC and ROC, we also eliminated 3, 5, and 4 patients respectively. 6.4 Correlation with Sevoflurane Concentration The correlation with the estimated effect-site sevoflurane concentration is evaluated for the spontaneous breathing (SB) period (table 6.2). The data number of SB is 5810 (23240 s) and the average duration for each patient is 750 ± 322 s. The overall indices ranking of Chapter 6. Results 42 Figure 6.12: Tracings and their faded error bars about the prediction probability (PK) values of the parameters and their standard errors in serial PK analysis for loss of con- sciousness (LOC). The baseline is one minute before LOC. NRR = non-rhythmic to rhythmic ratio; HF = high frequency power; LF = low frequency power; LHR = LF to HF ratio; BIS = Bispectral Index; HR = heart rate. (From ref.[69]) the PK analysis and Speareman rank correlation (R) are consistent. While BIS correlates with effect-site sevoflurane concentration best (SB: PK = 0.839, R = −0.836, p < 0.0001), the NRR index is second (PK = 0.736, R = −0.619, p < 0.0001). The tvLF is the best HRV indices ( PK = 0.669, r = −0.416, p < 0.0001). Compared to tvLF, the NRR index is significantly better ( p < 0.01). BIS surpasses NRR significantly in SB (p < 0.001). The NRR index is significantly better than HR (p < 0.0001). 6.5 Quantitative Results of Noxious Stimulation When observing the serial PK analysis for skin incision (Fig. 6.4), tvLF reaches the maximum amount (PK > 0.95) 30 seconds after skin incision. This shows that the tvLF is the best index of the skin incision, followed by the heart rate (PK < 0.85). Furthermore, tvLF is significantly better than BIS (PK < 0.55) and the NRR index (PK < 0.65). 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 -50 -25 0 25 50 75 100 PK time (s) Serial PK for LOC LHR NRR LF HR HF BIS Chapter 6. Results 43 Figure 6.13: Tracings and their faded error bars about the prediction probability (PK) values of the parameters and their standard errors in serial PK analysis for skin incision. The baseline is 5 seconds before skin incision. NRR = non-rhythmic to rhythmic ratio; HF = high frequency power; LF = low frequency power; LHR = LF to HF ratio; BIS = Bispectral Index; HR = heart rate. *P < 0.05, LF versus BIS and LF versus NRR. (From ref.[69]) In the serial PK analysis for endotracheal intubation (Fig. 6.16), I compared different indices calculated from multitaper reassignment spectrogram (HF, LF, LHR) with the in- dices calculated from STFT spectrogram (HFSTFT , LFSTFT , LHRSTFT ). This analysis signifies that the LF power and LHR provide the best results for the noxious stimulation of endotracheal intubation. Meanwhile, the heart rate is represented the second best and the BIS index is represented the worst. The indices that are calculated from multita- per reassignment spectrogram performs better than the corresponding indicators that are calculated from the STFT spectrogram. 6.6 Summary Several features on TF plane can be easily seen conveniently via the tvPS, including the "rhythmic-to-non-rhythmic phenomenon", "multiple component phenomenon", and 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 30 60 90 PK Time (s) Serial PK for Incision LF LHR HF HR NRR BIS * * * * * * * * Chapter 6. Results 44 Figure 6.14: Tracings and their faded error bars about the prediction probability (PK) values of the parameters and their standard errors in serial PK analysis for first motor movement reaction. The baseline is 3 minutes before first motor movement reaction. NRR = non-rhythmic to rhythmic ratio; HF = high frequency power; LF = low frequency power; LHR = LF to HF ratio; BIS = Bispectral Index; HR = heart rate. * P < 0.05, NRR versus BIS. (From ref.[69]) "LF surge phenomenon". The "rhythmic-to-non-rhythmic phenomenon" and "LF surge phenomenon" are apparent in the majority of my study cases. This universal consistency foreshadows the subsequent positive statistical results. The quantitative analysis proves that the NRR index indicates the first motor movement reaction better than other clinical indices[69]. BIS index is the best indiator for LOC and ROC. NRR index correlates with the concentration of sevoflurane. Furthermore, LF power indicates noxious stimulation best, and better than the heart rate. It is also shown the statistical benefit of the multitaper reassignment spectrogram over the classical STFT spectrogram. 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 -180 -150 -120 -90 -60 -30 0 30 60 90 120 150 PK time (s) Serial PK for First Reaction NRR BIS HR LHR LF HF * * * * * * Chapter 6. Results 45 Figure 6.15: Tracings and their faded error bars about the prediction probability (PK) values of the parameters and their standard errors in serial PK analysis for return of consciousness (ROC). The baseline is 3 minutes before ROC. NRR = non-rhythmic to rhythmic ratio; HF = high frequency power; LF = low frequency power; LHR = LF to HF ratio; BIS = Bispectral Index; HR = heart rate. * P < 0.05, BIS versus NRR. (From ref.[69]) 0.3 0.4 0.5 0.6 0.7 0.8 0.9 -180 -150 -120 -90 -60 -30 0 PK Time (s) Serial PK for ROC BIS HR NRR LF LHR HF * * * * * * * Chapter 6. Results 46 Figure 6.16: Tracings and their faded error bars about the prediction probability (PK) values of the parameters and their standard errors in serial PK analysis for endotra- cheal intubation. The baseline is 5 seconds before endotracheal intubation. HF = high frequency power calculated by multitaper reassigned spectrogram; LF = low frequency power calculated by multitaper reassigned spectrogram; LHR = LF to HF ratio calcu- lated by multitaper reassigned spectrogram; HF stft = high frequency power calculated by STFT spectrogram; LF stft = low frequency power calculated by STFT spectrogram; BIS = Bispectral Index; HR = heart rate. 0.4 0.5 0.6 0.7 0.8 0.9 0 25 50 75 100 PK Time (s) Serial PK for Intubation HF HF_stft LF LF_stft LHR LHR_stft BIS HR Chapter 6. Results 47 Table 6.1: Representative values of the Indices during important anesthetic events. Values are presented as median values (lower and upper quartiles). HF, LF, and LHR are expressed in common logarithm (10 base) of millisecond square. Heart rate is expressed in beat per minute. BIS = Bispectral Index; NRR = non-rhythmic to rhythmic ratio; HF = high frequency power; LF = low frequency power; LHR = LF to HF ratio; LOC = loss of consciousness; ROC = return of consciousness; T0= 60 s before LOC; T1= 60 s after LOC; T2= 5 s before incision; T3= 30s after incision; T4= 180 s before first reaction; T5= 5 s after first reaction; T6= 180 s before ROC; T7= 5 s after ROC. (From ref.[69]) LOC Skin incision First reaction ROC BIS NRR HF LF LHR T0 97 (94,98) -0.26 (-0.42, 0.21) 2.37 (2.05, 2.76) 1.84 (1.50, 2.27) -0.42 (-0.84, -0.27) T1 54 68) T2 T3 (42, 39 (36,44) 41 (37,44) T4 61 47) T5 (70, 73 (67,80) T6 80 83) T7 (77, 86 (81,94) 0.09 (-0.09, 0.31) 1.50 (0.97, 1.66) 1.37 (1.01, 1.60) -0.03 (-0.29, 0.19) -0.14 (-0.51, 0.02) 0.04 (-0.19, 0.21) 1.21 (0.94, 1.55) 0.84 (0.46, 0.94) -0.50 (0.46, 0.84) 1.42 (1.08, 1.81) 1.59 (1.25, 2.25) 0.24 (0.01, 0.50) -0.80 (-1.1, -0.38) 1.44 (1.09, 1.83) 0.09 (-0.37, 0.56) -1.33 (-1.67, -0.66) 0.47 (0.29, 0.61) 1.07 (0.78, 1.23) 0.79 (0.75, 1.31) 0.36 (0.12, 0.51) 1.11 (0.91, 1.40) 0.68 (0.01, 1.27) 0.43 (0.19, 0.62) 1.25 (0.93, 1.60) 0.91 (0.37, 1.26) -0.25 (-0.73, 0.06) -0.44 (-0.79, 0.10) -0.34 (-0.66, -0.12) HR (65, 74 91) 64 72) (58, (58, 65 72) (60, 70 76) (62, 71 80) (69, 79 85) 76 83) (69, 84 91) (75, Chapter 6. Results 48 Table 6.2: Prediction Probability and Spearman Rank Correlation for Indices vs. Es- timated Sevoflurane Effect-site Concentration in Controlled Ventilation and Spontaneous Breathing. (cid:63) p < 0.05 between non-rhythmic to rhythmic ratio (NRR) and other indices. Indices with prediction probability (PK) value smaller than 0.5 was corrected by 1 − PK before comparison. BIS = Bispectral Index; tvHF = high frequency power; tvLF = low frequency power; tvLHR = LF to HF ratio; HR = heart rate. (From ref.[69]) Controlled Ventilation PK (SE) Spearman rank correla- tion (95% CI) 0.716 (0.020) -0.575 (-0.661, -0.476) 0.670 (0.025) -0.467 (-0.583, -0.337) 0.479 (0.027)(cid:63) 0.073 (-0.069, 0.211)(cid:63) 0.582 (0.025) -0.233 (-0.359, -0.101)(cid:63) BIS NRR tvHF tvLF tvLHR 0.581 (0.025) -0.233 (-0.364, -0.096)(cid:63) HR 0.423 (0.024)(cid:63) 0.152 (0.034, 0.269)(cid:63) Spontaneous Breathing PK (SE) Spearman rank correla- tion (95% CI) BIS index 0.839 (0.014)(cid:63) -0.836 (-0.881, -0.778)(cid:63) NRR tvHF tvLF 0.736 (0.019) -0.619 (-0.688, -0.536) 0.489 (0.023)(cid:63) 0.001 (-0.117, 0.113)(cid:63) 0.669 (0.022)(cid:63) -0.416 (-0.506, -0.317)(cid:63) tvLHR 0.666 (0.022)(cid:63) -0.420 (-0.514, -0.319)(cid:63) HR 0.532 (0.022)(cid:63) 0.027 (-0.075, 0.133)(cid:63) Chapter 7 Implications to Anesthesiology and Physiology 7.1 Main Findings There are two important clinical results in regards to the NRR index. First, the NRR index is able to predict first reaction during emergency period. Second, the NRR index correlates with the sevoflurance concentration during the patient's spontaneous breathing. Even so, the NRR index is not the best predictor for LOC and ROC. These findings suggest that there is a distinct physiological interpretation of the anesthetic depth level when compared to the hypnosis measured by surface EEG. Furthermore, tvLF correlates well with skin incision and endotracheal intubation. This momentary information is currently unavailable to clinical anesthesiologists as both the IHR and tvPS are not implemented in current monitor. 7.2 Genuineness of NRR index The NRR index generally has a sharp increase nearly before the onset of the patient's first reaction. This then makes it highly unlikely that the motion artifacts interfere the NRR index. Furthermore, the observed NRR index performance in not confounded by the ventilation mode change. In particular, aside from the LOC, there are no transitions that occur from the mechanical ventilation to the spontaneous breathing in the data we have obtained. Also, there is no positive pressure ventilation for all the first reaction data. Furthermore, past literature on the association between "rhythmic-to-non-rhythmic" transition of the oscillatory pattern in RRI index and the different level of anesthesia has 49 Chapter 7. Implications 50 been mentioned but no quantitative works have been done. By using the classical power spectral analysis, Kato et al. has described that the high frequency component of RRI disperses when the patient awakens and concentrates when the patient is under anesthetics and controlled ventilation[2]. Based on the above facts, I am more than confident to state that the NRR is convincing in its distinct and novel information that it is capable of providing us. 7.3 Existing Findings The results we have obtained from the BIS (Fig.6.12, Table 6.2) are in agreement with previous studies showing that BIS is capable of monitoring awake status vs. anesthesia, the first reaction and the decrease of anesthetic gas concentration[5, 8, 10]. Our BIS results (Fig.6.13, Fig.6.16) also agree with past studies that provide evidience that BIS or other EEG-derived indices cannot indicate the response to noxious stimulation[6, 7, 9]. The results from our analysis (Table 6.1) are also consistent with previous reported associ- ation between anesthetic depth and traditional HRV parameters: HRV is decreased during general anesthesia and increased during recovery[2, 46, 70, 71]. Lastly, our results are also consistent with the reported finding that combining classical HRV indices with BIS does not outperform BIS in discriminating awake from asleep during anesthetic induction[72]. It has been reported that classical HRV parameters do not outperform heart rate when predicting noxious stimulation[4, 73]. Contrarily, our results demonstrate that tvLF pre- dicts skin incision better than HR. It seems that the noxious stimuli of the skin incision elicited a momentary increase of sympathetic activity, which leads to this finding. Al- though the study is not designed for noxious stimulation, the finding that tvLF correlates with skin incision shows the technical advantage of our approach and its potential as an index for noxious stimulation. 7.4 Physiologic Interpretation of NRR The association between NRR and anesthetic depth can be partially explained by the cardiopulmonary coupling[14] and the respiratory physiology. When the coupling effect is higher, the respiration pattern is reflected in the RRI index. The respiratory mecha- nism also contributes to the differential influence that occur from anesthetics. The neural respiratory control comprises of two systems, the involuntary automatic control system and the voluntary control system. The involuntary automatic control system is mainly Chapter 7. Implications 51 controlled by the respiratory center in the pontomedullary area. The voluntary control sys- tem is mainly controlled by the forebrain[25]. These two systems are also distinct in their neural anatomy and their functions. Respiratory efferent signals from these two systems compete with each other and are integrated at the spinal level to control the respiratory motor neuron. However, during spontaneous respiration, studies have also shown that the involuntary automatic respiratory pattern generated in the pontomedullary center[25], specifically the preBotzinger complex[26 -- 28], is rhythmic. On the other hand, the breathing pattern of the voluntary respiratory motor control is non-rhythmic and involves the cortical processing and thalamic integration which responds to the peripheral inputs and the descending inputs[25, 30 -- 32]. The thalamus also ac- tively participates in different motor functions, including respiration[31, 32], speech[33] and cough[34] to name a few[74]. Our NRR index findings suggest that the non-rhythmic respiratory activity involves the nearly entire cerebral regions is more susceptible to sevoflu- rane than the rhythmic respiration generated in the medulla. This suggestion is supported by literature. Guedel made the classical observation on respiratory pattern[1]. Bimer et al denoted that the respiratory irregularities accompany electroencephalographic arousal reaction in spontaneous breathing[75]. Also, Studies have also shown that the respiratory activity of preBotzinger complex is less depressed by sevoflurane[36, 37], and the effect of volatile anesthetics is less prominent to the brainstem compared to the cortical regions that are more abundant in synaptic transmissions.[37]. The ability for the NRR index to predict first reaction suggests a connection between the "rhythmic-to-non-rhythmic" phenomenon and motor reactions. Antognini et al. demon- strated that immobility to surgical stimulation by volatile anesthetics in goats is mainly modulated by the spinal cord, which also indirectly affects the thalamic response to nox- ious stimulation[12, 13]. Another study by Velly et al. also presented significant results[7]. He recorded the human subcortical electrophysiological activity and demonstrated that subcortical activity (possible the thalamic activicy) predicts suppression of movement to noxious stimuli, but not changes of consciousness, whereas cortical EEG predicts loss of consciousness, but not motor suppression. Based on the above, NRR might reflect the subcortical activity since it better reveals information about motor reaction, but not consciousness, compared with BIS. Since immobility to noxious stimulation is similar to, although not the same as the first reaction under surgical wound pain in the present study, this relationship suggests the possible role of the thalamus in "rhythmic-to-non-rhythmic" phenomenon. Besides the NRR index, BIS also responded to first reaction in our serial PK analysis, which is in agreement with other studies[10]. Therefore, it is possible that a cortical mechanism is also related to the first reaction. Chapter 7. Implications 52 From these results, it is possible to see that sevoflurane may affect non-rhythmic respiration level of the spinal cord, subcortical supraspinal regions, or the cortex in the human brain. Although we are currently unable to pinpoint the exact anatomic location, it is likely that the subcortial supraspinal area, possibly the thalamus, is the most plausible region. The above statement is sound as the NRR index generates different results in the serial PK analysis and is also closely connected to the human body's respiration responses. Thus, we hypothesize that the "rhythmic-to-non-rhythmic" phenomenon in IHR exhibits the central respiratory activity via either central or peripheral mechanisms[14]. The IHR also reflects the final integration of this "rhythmic-to-non-rhythmic" respiratory activity. Although more evidence is necessary to further clarify these hypotheses, we propose to use the NRR quantification methodology as a potential tool to evaluate the depth of anesthesia that differs from EEG-based monitoring. 7.5 Clinical Application of the NRR Index The first strength of this study is a new quantitative approach, referred to as the NRR index, to analyze "rhythmic-to-non-rhythmic" phenomenon. The NRR index has the po- tential to reflect different levels of anesthesia. Literature partially supports that there is an underlying physiological mechanism. Second, the signal processing technique, mul- titaper Synchrosqueezing transform, has been well studied in the literature, so we have adequate theoretical support to resolve the potential limitations of the PS. The momentary dynamics in IHR can be captured by the tvPS, which leads to the NRR index. 7.6 Potential Index in Noxious Stimulation Although we did not design the study for noxious stimulation, from the data analysis result, we found that the indices we apply are potential in detecting pain reaction under anesthesia. In particular, the tvLF index correlates with the noxious stimulation better than heart rate as an additional finding. More study will need to be conducted on the tvLF to provide sound results. It can also be helpful to compare this data with other well known indices like the Surgical Stress Index ® (SSI, GE healthcare) and Analgesia Nociception Index ® (ANI, Metrodoloris), which have been used on monitoring surgical stress [11]. SSI is derived from the pulse wave amplitude and pulse beat interval of the photoplethysmography whereas ANI indicates the parasympathetic tone derived from ECG signal. It is possible that various indices can provide complementary information regarding noxious stimulation that is gathered from different modalities and concepts. Chapter 7. Implications 53 NRR index measures the rhythmic-to-non-rhythmic transition and quantifies a new kind of information different from the above two instruments in both physiologic and mathematical senses. Further investigations may be necessary to understand the clinical value and application of tvPS. 7.7 Summary In conclusion, we are able to extract hidden information regarding the anesthetic levels from the routine ECG monitor. The quantitative results of the NRR index supports our initial hypothesis that the "rhythmic-to-non-rhythmic" transition correlates with motor reactions during emergence period earlier than BIS index, and correlates with sevoflu- rance concentration. In addition, the potential of the time-varying HRV indices provides observations of the relationship between tvLF index and the noxious stimulation. The notion of these dynamics is rooted in the recently developed signal processing technique. Without tvPS, the above quantification cannot be easily achieved. Overall, my clinical study suggests that the ECG signal contains complementary information to the EEG-based depth-of-anesthesia index. Chapter 8 Real-time Processing Using the Blending Operator 8.1 Real-time Processing From the previous chapters, my study has concluded that the IHR signal during anesthesia contains a wealth of information when viewing with tvP S. It is possible to derive the IHR signal from ECG, the waveform of pulse oximetry, and also the invasive intra-arterial pressure waveform. Although we mainly use ECG signal in the present project, all possible sources of IHR are mandatory or frequently used functions in modern standard anesthesia monitors. A practical question is how do we apply the NRR index proposed in the present study to individual patients. The processing steps of NRR index starts from the identification of each heart beat from ECG signal. Next, an interpolation is used to convert the irregularly-sampled heart beats data into regularly-sampled IHR data. Next, the calculation of the tvP S is based on the multitaper Synchrosqueezing tranform. And lastly the NRR index is computed based on the tvP S. Thus, the positive statistical results in previous chapter, are all obtained from off-line calculations. In anesthesia practices, the main purpose of anesthesia monitor is to provide real time monitoring of the patient's responses in order to provide timely responses. This purpose then turns ordinary measurements like heart rate, blood oxygen saturation, and blood pressure into vital information during anesthesia. The real-time processing of NRR in- dex is necessary if we would like to "steer" the anesthetic depth of patients based on their instantaneous information from IHR. The potential benefits include adequate dose 54 Chapter 8. Real-time Processing 55 of anesthetics, possibility to divert critical situation, reduce stress response, and improve long-term welfare. In this chapter, my main focus is the real-time processing from the irregular heart beat sample to the tvP S. The goal is to interpolate the irregular data samples in real time. Meanwhile, the polynomials property of the interpolation also serves as the vanishing- moment and minimum-supported wavelets. 8.2 Technique Obstacles for Real-time Processing Due to the irregular interval of each heartbeat, the heartbeat data is irregularly sampled and this irregularity is unavoidable. Almost all continuous information are digitized in equally-spaced samples in time or in space before further analysis for convenience, and for efficiency. However, some situations provide non-uniform sampling and could be either unavoidable or intended[76, 77]. In order to derive the continuous instantaneous heart rate from the distinct heartbeats (Fig.2.1) is an example of the rich IHR information that entirely relies on the irregularity of the heartbeats. Hence, the initial challenge is the real-time processing of the non-uniform data samples. When analyzing data the scope of continuous wavelet transform (CWT), the theoretical background of Synchrosqueezing transform presents the optimal performance when us- ing compactly supported windows in frequency domain[44]. That means we need whole data converted in frequency domain for subsequent calculation. This is inefficient and also highly unlikely for real-time data processing. Moreover, for real-time processing, an wavelet transform (or STFT) better has a short, and compact support window in time. This type of window may hamper the benefit of wavelet transform: an adequate vanishing moment can eliminate trend component in terms of corresponding polynomial order to preserve the oscillatory components we desire more efficiently. A possible workaround is by cutting down a segment of data, windowing it, and then choosing a suboptimal wavelet function empirically. Next, we would observe the effect of suboptimal window and then adjust the parameters empirically until we finally accept the suboptimal effect in real-time display. There is a better solution to consider before we proceed with the above workaround. The solution is a special spline interpolation, which has the interpolating ability to handle the unequally-spaced heart beat sample, and the polynomial property to match the vanishing moment property of wavelet transform. The resulting time-scale analysis is compactly supported in time and provides better performance due to the vanishing moment property. Chapter 8. Real-time Processing 56 Based on this scheme, the real-time tvPS can help remove the above mentioned technical obstacles. 8.3 B-spline We start from a brief introduction of the spline function. For digital samples g(tn) are sampled at time points t : a = t0 < t1 < t2 < . . . , (8.1) where {tk} may be irregular, or non-uniform signal, which means that we allow tk+1− tk (cid:54)= tj+1 − tj for k (cid:54)= j. Let m ≥ 3 be any desired integer and Πm−1 denote the space of all polynomials of degree less than m. The spline space is then defined as: s : s−m+1 = s−m+2 = . . . = a = s0 < s1 < s2 < . . . , (8.2) and the notation Ss,m := Ss,m[a,∞) denotes the spline space of order m with the knot sequence s. To formulate a locally supported basis of Ss,m, we now introduce the truncated powers xm−1 + := (max{0, x})m−1. The truncated power functions (sk − t)m−1 + , k = 0, 1, . . ., are in Ss,m. Since (xk − t)m−1 + , k ≥ 1, have global support, we apply the m-th order divided differences to form locally supported functions. The divided differences are [ u, . . . , u ]f := f (l)(u) l! if there are l + 1 entries in [u, . . . , u], and [ u0, . . . , un ]f := [ u1, . . . , un ]f − [ u0, . . . , un−1 ]f un − u0 if u0 ≤ u1 ≤ . . . ≤ un and un > u0. Now we consider the knot sequence s of the spline space Ss,m. The normalized B-splines[78] are the divided difference of the truncated power (sk − t)m−1 + , namely, Nm,k(t) = Ns,m,k(t) = (sm+k − sk)[ sk, . . . , sm+k ](· − t)m−1 + , (8.3) for k = −m + 1, . . . , 0, 1, 2, . . .. Chapter 8. Real-time Processing 57 We would apply the recursive formula to implement Nm,k [52, page 143 (6.6.12a)]. The B-splines are a locally supported. However, if we choose the irregularly-spaced sequence t of time positions in (8.1) as the knot sequences s in (8.2) by attaching t−m+1 = . . . = t0 = a to t as in (8.2), it can be difficult to compute, or the computation cost will be higher for the non-uniform {tj} and large values of n. 8.4 Quasi-interpolation On the other hand, de Boor and Fix [79] proposed a spline representer f (t), which ap- proximates the data instead of providing exact interpolating the target data function g(t) at t = tj, j = 0, 1, . . . , n (that is, if f (tj) (cid:54)= g(tj) is allowed). This quasi-interpolation has a "polynomial preservation" property: g(t) = p(t) ∈ Πm−1. This polynomial preservation is important to improve the performance of the continuous wavelet transform (CWT), and hence the Synchrosqueezing transform (SST), to reveal the oscillatory components. Specifically, if the analysis wavelet is compactly supported by the spline of order m, then the m-th order vanishing moment annihilates the m-th order Taylor polynomial expansion of g(t). This then facilitates the trend removal of the the signal g(t) as a polynomial. However, the quasi-interpolation scheme de Boor proposed [78, 79] requires derivative data values of g(t). Others [80, 81] proposed to replace the derivatives of g(t) by divided differences of {g(ti)}. However, these quasi-interpolations[82] do not address the real-time issue at stake. In addition, quasi-interpolation introduced the error g(ti) − f (ti) for i = 0, 1, . . . , n. Chui et al. proposed a real-time quasi-interpolation[83]. Chui and his college also intro- duced a "local interpolation"[84] to correct the error of "quasi-interpolation". The real-time quasi-interpolant [83] is introduced as follow: for each k = 0, 1, . . ., consider the Vandermonde determinant D(tk, . . . , tk+m+1) := det  1 1 tk ... tm−1 k tk+1 ... tm−1 k+1 . . . . . . ... . . . 1 tk+m−1 ... tm−1 k+m−1  , and the determinant D(tk, . . . , tk+j−1, ξj, tk+j+1, . . . , tk + m− 1) obtained by replacing the (j + 1)-st column in the definition of D(tk, . . . , tk+m+1) by the column vector ξj := [ ξ0(j, m), . . . , ξm−1(j, m) ]T , Chapter 8. Real-time Processing 58 where ξ0(j, m) = 1 and ξi(j, m) = σi(tj+1, . . . , tj+m−1) , (cid:32) (cid:33) for i = 1, . . . , m − 1 with σi(r1, . . . , rm−1) being the classical symmetric functions defined by σ0(r1, . . . , rm−1) = 1 and for i = 1, . . . , m − 1, σi(r1, . . . , rm−1) = 1≤l1<...<li≤m−1 rl1 . . . rli. m − 1 i (cid:88) 2m−2(cid:88) l=m−1 (cid:88) Using the determinants introduced above, we apply the below spline coefficients ak,l := D(tk, . . . , tk+j−1, ξj, tk+j+1, . . . , tk+m−1) D(tk, . . . , tk+m−1) , to formulate the compactly supported spline function Mt,m,k(t) := ak,l−m+1Nt,m,k+l−m+1(t), with suppMt,m,k = [tk−m+1, tk+m]. These basis functions provide a real-time implementa- tion of the quasi-interpolation operator (Qmg)(t) = g(tk)Mt,m,k(t). (8.4) k The quasi-interpolation operator Qm (8.4) possesses the polynomial preservation property. For any m ≥ 1, (Qmp)(t) = p(t), for all t ≥ a and for all p ∈ Πm−1, provided that the summation (8.4) account for all non- negative integers k = 0, 1, . . .. Furthermore, in view of the support of Mt,m,k, it follows that v+m−2(cid:88) p(tk)Mt,m,k(t) = p(t), t ∈ [tu, tv] for all p ∈ Πm−1. k=v−m+1 This local polynomial preservation property allows the CWT, and hence the SST, to annihilate the (m − 1)-th degree Taylor polynomial approximation of the signal at tj, where u < j < v. Chapter 8. Real-time Processing 59 8.5 Local Interpolation Operator The local interpolation operator, denoted by Rm, satisfies the interpolation property. To define Rm, we will insert knots to t by considering a new knot sequence s ⊃ t (8.2). For simplicity, we will introduce even order variable m : For even m ≥ 4, we set smk/2 = tk, k = 0, 1, 2, . . . . That is, we insert (m/2− 1) knots in between two consecutive knots in t. For convenience, we will choose the new knots to be equally spaced between every pair of two consecutive knots. Fix even m ≥ 4. Let Ns,m,j be the m-th order B-spline with knot sequence s. Then the completely local spline basis function can be defined by Ls,m,j(t) := Ns,m,m(j−1)/2(t) Ns,m,m(j−1)/2(tj) . Since tj = smj/2 is the "centered" knot and we have suppNs,m,m(j−1)/2 = [ sm(j−1)/2, sm(j+1)/2 ] = [ tj−1, tj+1 ],  Ls,m,j(tj) = 1 suppLs,m,j = [ tj−1, tj+1 ]. (8.5) It is clear that Ls,m,j(tk) = δj−k, where the Kronecker delta notation is used (8.5). The above preparation provides a real-time implementation of the local interpolation op- erator, which satisfies the interpolation property (8.5). Fix m ≥ 3. For a given function g ∈ C(R), the local interpolation operator Rm is defined as (cid:88) (Rmg)(t) := g(tk)Ls,m,k(t). (8.6) k 8.6 Blending Operator We are now ready to apply (8.4) to obtain the blending operator, denoted by Rm ⊕ Qm. Chapter 8. Real-time Processing 60 Fix m ≥ 3 and g ∈ C(R). The blending operator is defined as Pm := Rm ⊕ Qm, where Rm ⊕ Qm := Qm + Rm(I − Qm) = Qm + Rm − RmQm, and I is the identity operator. In particular, we have (cid:88) (cid:88) (cid:2)g(tk) −(cid:88) g(tj)Mt,m,j(tk)(cid:3)Ls,m,k(t). (8.7) (Pmg)(t) := g(tk)Mt,m,k(t) + k k j We remark that in the definition of Pm, the two operators Rm and Qm are not commutative. Let us summarize the two key properties of the blending operator in the following theorem. 1. The blending operator Pm possesses both the polynomial preservation property of Qm and the interpolatory property of Rm. 2. The error of spline interpolation by the blending operator is both small and bounded. In conclusion, the blending operator, as a local spline interpolation operator, achieves the optimal interpolation error rate when compared with the traditional spline interpolation operator. In addition, the error depends only on the local data profile, which allows real-time implementation with optimal error rate. 8.7 Real-time tvPS We provide a real-time computational algorithm to compute f (t) for the upcoming data samples g(t0), g(t1), . . . etc. For convenience, we only consider even order m. The formu- lation for the odd order is similar but slightly more complicated. The blending operator leads to the real-time spline interpolation. For real-time issue, the wavelets have to be minimum supported in time. We can design the wavelets with the same number of vanishing moments as the spline. Meanwhile the wavelets have minimum support and maximum order of vanishing moments. Let m, n ≥ 1 be arbitrary integers, and x an arbitrary knot sequence. The spline basis functions is then as follows: ψx,m;n,k(x) := N (n) x,m+n,k(x), k ∈ Z, (8.8) Chapter 8. Real-time Processing 61 Algorithm 1 Real-time implementation of the blending operator First, pre-compute the B-spline values Ns,m,l(tj) =: nl,j. Then for each k, since Mt,m,k ∈ Ss,m, there exists a finite sequence {bk,l} in the formulation of Mt,m,k(t) = bk,lNs,m,l(t). Also pre-compute dk,j = Mt,m,k(tj). Now, while the data sequence {g(tk)} is acquired, compute (cid:88) l gl = bk,lg(tk) (cid:88) g(tl) −(cid:80) k and simultaneously compute and then up-sample {g∗ g∗ l = j dl,jg(tj) , nm(l−1)/2, l l } by m(l − 1)/2; that is, set and g# l = 0. m(k−1)/2 = g∗ g# k, n(cid:88) f (t) = l=−m+1 (gl − g# l )Ns,m,l(t) Then, we have an on-line computational scheme for the quasi-interpolation spline inter- polation: for increasing number of samples from g(tn) to g(tn+1), . . .; and this can be implemented for real-time D/A conversion. to be called VM wavelets on x, where Nx,m+n,k(x) is defined in (8.3), which satisfies the moment conditions  (cid:90) ∞ (cid:90) ∞ −∞ −∞ xlψx,m;n,k(x)dx = 0, l = 0, . . . , n − 1 xnψx,m;n,k(x)dx (cid:54)= 0. (8.9) That means the arbitrary chosen n in the interpolator leads to the vanishing moment order of the wavelets if we use this spline basis as the wavelet function. The wavelet transform can successfully eliminate the n order polynomial as a trend component. Let m, n ≥ 1 be arbitrary integers. Then the derivative of the wavelet ψx,m;n,k on an arbitrary knot sequence x, is given by ψ(cid:48) x,m;n,k(x) = ψx,m−1;n+1,k(x). (8.10) Chapter 8. Real-time Processing 62 This property permits us to conveniently calculate the derivative of the wavelet, which is required to calculate the frequency reassignment in Synchrosqueezing transform.(see equation 4.3) Then, we can proceed to the Synchrosqueezed CWT to obtain the real-time time-frequency analysis as real-time tvPS. Using this tvPS, we can apply the research in previous chap- ters to quantify the "rhythmic-to-non-rhythmic" phenomenon in real-time. Using the clinical data to "emulate" the real-time calculation of NRR index, we obtained a compa- rable performance of the correlation with sevoflurane concentration during spontaneous respiration: The PK value of real-time NRR index is 0.711 ± 0.021 (PK of off-line NRR index:0.732 ± 0.018)(6.2). Chapter 9 Conclusion 9.1 Research Findings Thanks to the advancement of science in anesthesiology and neural physiology, I can pro- pose a theory and modeling to investigate the "rhythmic-to-non-rhythmic" phenomenon. Furthermore, because of the advancement of signal processing technique, I am able to use multitaper Synchrosqueezing technique to compute the NRR index. Anesthesia creates dynamic changes to the human body, and due to the time-varying ability and good TF resolution, the NRR index is a particularly suitable tool reflecting the constant changes during anesthesia[43, 61]. The numeric result from clinical database is positive [69]. NRR index performs better and faster than BIS index when detecting motor movement reaction. This also supports the fact that different brain regions mediating various functions, with differential susceptibility to anesthetics may require multi-modal assessment of these anesthetic effects. NRR index is therefore able to provide additional value. In addition, LF index provides good results as an indicator for noxious stimulation, for both skin incision and endotracheal intubation. ECG waveform is standard information in current anesthesia, while the IHR is not. My study shows that these new indices derived from ECG waveform can provide additional information that could be beneficial to patients undergoing anesthesia in the future. 9.2 Accomplishments 1. The adaptive harmonic model is a method of modeling the rhythmicity of the "rhythmic- to-non-rhythmic" phenomenon 63 Chapter 9. Conclusion 64 Figure 9.1: The diagram represents the overall framework from the human brain to signal analysis. 2. A time-varying power spectrum that is based on multitapering Synchrosqueezed spectrogram. 3. Methodology of the NRR index 4. A methodology of serial PK analysis was developed to evaluate the dynamic perfor- mance of anesthetic depth index as time-varying PK value. 5. Positive clinical value of the NRR index was obtained from a clinical database 6. Improved performance of classical HRV parameters was obtained using tvP S 7. A real-time scheme that provides real-time tvP S and real-time NRR index. 9.3 Future Directions The results of my project indicates several future research possibilities. In clinical anesthesiology, more studies should be conducted to answer questions regard- ing the value of NRR index in different kinds of anesthetics (i.e. propofol, ketamine, dexmedetomidine or xenon), in different types of patient (i.e. children, lying-in women, Chapter 9. Conclusion 65 elderly patient, or patient with severe illness). Since the NRR index may indicate the activity of subcortical areas, this poses questions on if the application of the NRR index can lead to less stress response and better long-term outcomes, and if it is possible, how can we obtain those results. On the other hand, the IHR is derived from the ECG signal, I wonder whether it is possible to calculate the NRR index from photoplethsmography or intra-arterial pressure waveform. It also needs more clinical study to better understand the clinical role of NRR index and the tvLF index. It is possible that a thorough understanding of the basic mechanism can lead to better clinical benefit. The origin of "rhythmic-to-non-rhythmic" phenomenon should be in the brain. However, the neural mechanism has not been addressed according to my best knowl- edge. For example, where exactly is the anatomical location that creates this phenomenon? What could cause the multiple component phenomenon as previously mentioned? Lastly, how do the neurons coordinate with one another during this phenomenon? It is foreseeable that the answers of these questions can directly contribute to the brain science. Thus, the neurophysiological study based on the animal model is a potential future research. The signal processing technique directly contributes to the performance of these clinical indices. The current result of tvPS is helpful in revealing various features in the TF plane, but there are still a lot of room for improvements. The time-frequency analysis is based mostly on the Fourier transform. I am interested in whether it is possible to discard the idea of frequency in order to have a better understanding of the signal's features. The real- time calculation and display of the clinical indices provides benefit to individual patients. Technique developments in real-time processing is imperative. 9.4 Summary My study focuses on the qualifications of the "rhythmic-to-non-rhythmic" phenomenon. I have verified that my proposed model and quantitative index through using clinical database. The positive results support my proposal on using the adaptive harmonic model and time-varying power spectrum based on the multitaper Synchrosqueezing transform. The NRR index also presents unique values in clinical anesthesia. The achievements in this thesis fill a gap in the fields of medicine and provide a new perspective on many of the issues that could be understood in the future. Bibliography [1] Guedel AE. Inhalational Anesthesia: A Fundamental Guide. New York, Macmillan, 1937. [2] Kato M, Komatsu T, Kimura T, Sugiyama F, Nakashima K, and Shimada Y. Spectral analysis of heart rate variability during isoflurane anesthesia. Anesthesiology, 77(4): 669 -- 674, 1992. [3] Miller RD, Eriksson LI, Fleisher L, Wiener-Kronish JP, and Cohen NH. Miller's anesthesia. Elsevier Health Sciences, 2014. [4] Seitsonen ERJ, Korhonen IKJ, Van Gils MJ, Huiku M, Lotjonen JMP, Korttila KT, and Yli-Hankala AM. EEG spectral entropy, heart rate, photoplethysmography and motor responses to skin incision during sevoflurane anaesthesia. Acta Anaesthesiolog- ica Scandinavica, 49(3):284 -- 292, 2005. [5] Vakkuri A, Yli-Hankala A, Talja P, Mustola S, Tolvanen-Laakso H, Sampson Tl, and Viertio-Oja H. Time-frequency balanced spectral entropy as a measure of anesthetic drug effect in central nervous system during sevoflurane, propofol, and thiopental anesthesia. Acta Anaesthesiologica Scandinavica, 48(2):145 -- 153, 2004. [6] Storm H, Støen R, Klepstad P, Skorpen F, Qvigstad E, and Raeder J. Nociceptive stimuli responses at different levels of general anaesthesia and genetic variability. Acta Anaesthesiologica Scandinavica, 57(1):89 -- 99, 2013. [7] Velly LJ, Rey MF, Bruder NJ, Gouvitsos FA, Witjas T, Regis JM, Peragut JC, and Gouin FM. Differential dynamic of action on cortical and subcortical structures of anesthetic agents during induction of anesthesia. Anesthesiology, 107(2):202 -- 212, 2007. [8] Soehle M, Ellerkmann RK, Grube M, Kuech M, Wirz S, Hoeft A, and Bruhn J. Comparison between bispectral index and patient state index as measures of the electroencephalographic effects of sevoflurane. Anesthesiology, 109(5):799 -- 805, 2008. 66 Bibliography 67 [9] Ekman A, Brudin L, and Sandin R. A comparison of bispectral index and rapidly extracted auditory evoked potentials index responses to noxious stimulation during sevoflurane anesthesia. Anesthesia & Analgesia, 99(4):1141 -- 1146, 2004. [10] Schmidt GN, Bischoff P, Standl T, Hellstern A, Teuber O, and Schulte Esch J. Com- parative evaluation of the datex-ohmeda s/5 entropy module and the bispectral index (r) monitor during propofol-remifentanil anesthesia. Anesthesiology, 101(6):1283 -- 1290, 2004. [11] Wennervirta J, Hynynen M, Koivusalo AM, Uutela K, Huiku M, and Vakkuri A. Sur- gical stress index as a measure of nociception/antinociception balance during general anesthesia. Acta anaesthesiologica Scandinavica, 52(8):1038 -- 1045, 2008. [12] Antognini JF and Schwartz K. Exaggerated anesthetic requirements in the preferen- tially anesthetized brain. Anesthesiology, 79(6):1244 -- 1249, 1993. [13] Antognini JF, Wang XW, and Carstens E. Isoflurane action in the spinal cord blunts electroencephalographic and thalamic -- reticular formation responses to noxious stim- ulation in goats. Anesthesiology, 92(2):559, 2000. [14] Eckberg DL. Point: counterpoint: respiratory sinus arrhythmia is due to a cen- tral mechanism vs. respiratory sinus arrhythmia is due to the baroreflex mechanism. Journal of applied physiology, 106(5):1740 -- 1742, 2009. [15] Poyhonen M, Syvaoja S, Hartikainen J, Ruokonen E, and Takala J. The effect of carbon dioxide, respiratory rate and tidal volume on human heart rate variability. Acta Anaesthesiologica Scandinavica, 48(1):93 -- 101, 2004. [16] Hall JE. Guyton and Hall Textbook of Medical Physiology: Enhanced E-book. Elsevier Health Sciences, 2010. [17] Detsch O, Kochs E, Siemers M, Bromm B, and Vahle-Hinz C. Differential effects of isoflurane on excitatory and inhibitory synaptic inputs to thalamic neurones in vivo. British Journal of Anaesthesia, 89(2):294 -- 300, 2002. [18] Avidan MS, Jacobsohn E, Glick D, Burnside BA, Zhang L, Villafranca A, Karl L, Kamal S, Torres B, O'Connor M, Evers AS, Gradwohl S, Lin N, Palanca BJ, and Mashour GA. Prevention of intraoperative awareness in a high-risk surgical popula- tion. New England Journal of Medicine, 365(7):591 -- 600, 2011. [19] Monk TG, Saini V, Weldon BC, and Sigl JC. Anesthetic management and one-year mortality after noncardiac surgery. Anesthesia & Analgesia, 100(1):4 -- 10, 2005. Bibliography 68 [20] Melamed R, Bar-Yosef S, Shakhar G, Shakhar K, and Ben-Eliyahu S. Suppression of natural killer cell activity and promotion of tumor metastasis by ketamine, thiopen- tal, and halothane, but not by propofol: mediating mechanisms and prophylactic measures. Anesthesia & Analgesia, 97(5):1331 -- 1339, 2003. [21] Anand KJS and Hickey PR. Halothane -- morphine compared with high-dose sufentanil for anesthesia and postoperative analgesia in neonatal cardiac surgery. New England Journal of Medicine, 326(1):1 -- 9, 1992. [22] Wallace AW, Galindez D, Salahieh A, Layug EL, Lazo EA, Haratonik KA, Boisvert DM, and Kardatzke D. Effect of clonidine on cardiovascular morbidity and mortality after noncardiac surgery. Anesthesiology, 101(2):284 -- 293, 2004. [23] Mangano DT, Layug EL, Wallace A, and Tateo I. Effect of atenolol on mortality and cardiovascular morbidity after noncardiac surgery. New England Journal of Medicine, 335(23):1713 -- 1721, 1996. [24] Bardram L, Funch-Jensen P, Jensen P, Kehlet H, and Crawford ME. Recovery after laparoscopic colonic surgery with epidural analgesia, and early oral nutrition and mobilisation. The Lancet, 345(8952):763 -- 764, 1995. [25] Mitchell RA and Berger AJ. Neural regulation of respiration. The American review of respiratory disease, 111(2):206 -- 224, 1975. [26] Elsen FP and Ramirez JM. Postnatal development differentially affects voltage- activated calcium currents in respiratory rhythmic versus nonrhythmic neurons of the pre-botzinger complex. Journal of neurophysiology, 94(2):1423 -- 1431, 2005. [27] Rekling JC and Feldman JL. Prebotzinger complex and pacemaker neurons: hypothe- sized site and kernel for respiratory rhythm generation. Annual Review of Physiology, 60(1):385 -- 405, 1998. [28] Ramirez JM, Quellmalz UJA, Wilken B, and Richter DW. The hypoxic response of neurones within the in vitro mammalian respiratory network. The Journal of Physiology, 507(2):571 -- 582, 1998. [29] Shea SA. Behavioural and arousal-related influences on breathing in humans. Exper- imental physiology, 81(1):1 -- 26, 1996. [30] McFarland NR and Haber SN. Thalamic relay nuclei of the basal ganglia form both reciprocal and nonreciprocal cortical connections, linking multiple frontal cortical areas. The Journal of neuroscience, 22(18):8117 -- 8132, 2002. [31] Ramsay SC, Adams L, Murphy K, Corfield DR, Grootoonk S, Bailey DL, Frackowiak RS, and Guz A. Regional cerebral blood flow during volitional expiration in man: Bibliography 69 a comparison with volitional inspiration. The Journal of physiology, 461(1):85 -- 101, 1993. [32] Mckay LC, Evans KC, Frackowiak RSJ, and Corfield DR. Neural correlates of volun- tary breathing in humans. Journal of Applied Physiology, 95(3):1170 -- 1178, 2003. [33] Murphy K, Corfield DR, Guz A, Fink GR, Wise RJS, Harrison J, and Adams L. Cerebral areas associated with motor control of speech in humans. Journal of Applied Physiology, 83(5):1438 -- 1447, 1997. [34] Simonyan K, Saad ZS, Loucks TMJ, Poletto CJ, and Ludlow CL. Functional neu- roanatomy of human voluntary cough and sniff production. Neuroimage, 37(2):401 -- 409, 2007. [35] Sloan TB. Anesthetic effects on electrophysiologic recordings. Journal of clinical neurophysiology, 15(3):217 -- 226, 1998. [36] Kuribayashi J, Sakuraba S, Kashiwagi M, Hatori E, Tsujita M, Hosokawa Y, Takeda J, and Kuwana S. Neural mechanisms of sevoflurane-induced respiratory depression in newborn rats. Anesthesiology, 109(2):233 -- 242, 2008. [37] Takita K and Morimoto Y. Effects of sevoflurane on respiratory rhythm oscillators in the medulla oblongata. Respiratory physiology & neurobiology, 173(1):86 -- 94, 2010. [38] Fink BR, Hanks EC, Ngai SH, and Papper EM. Central regulation of respiration during anesthesia and wakefulness. Annals of the New York Academy of Sciences, 109(2):892 -- 900, 1963. [39] Moody GB, Mark RG, Zoccola A, and Mantero S. Derivation of respiratory signals from multi-lead ECGs. Computers in cardiology, 12:113 -- 116, 1985. [40] Houtveen JH, Groot PF, and de Geus EJ. Validation of the thoracic impedance derived respiratory signal using multilevel analysis. International Journal of Psy- chophysiology, 59(2):97 -- 106, 2006. [41] Flaherty JT, Blumenschein SD, Alexander AW, Gentzler RD, Gallie TM, Boineau JP, and Spach MS. Influence of respiration on recording cardiac potentials: Isopo- tential surface-mapping and vectorcardiographic studies∗. The American journal of cardiology, 20(1):21 -- 28, 1967. [42] Task Force of the European Society of Cardiology, Task Force of the European Soci- ety of Cardiology, et al. the north american society of pacing and electrophysiology. heart rate variability: standards of measurement, physiological interpretation and clinical use. Circulation, 93(5):1043 -- 1065, 1996. Bibliography 70 [43] Lin YT, Hseu SS, Yien HW, and Tsao J. Analyzing autonomic activity in electrocar- diography about general anesthesia by spectrogram with multitaper time-frequency reassignment. In Biomedical Engineering and Informatics (BMEI), 2011 4th Inter- national Conference on, volume 2, pages 630 -- 634. IEEE, 2011. [44] Daubechies I, Lu J, and Wu HT. Synchrosqueezed wavelet transforms: an empirical mode decomposition-like tool. Applied and computational harmonic analysis, 30(2): 243 -- 261, 2011. [45] Chen YC, Cheng MY, and Wu HT. Non-parametric and adaptive modelling of dy- namic periodicity and trend with heteroscedastic and dependent errors. Journal of the Royal Statistical Society: Series B (Statistical Methodology), 76(3):651 -- 682, 2014. [46] Huang HH, Chan HL, Lin PL, Wu CP, and Huang CH. Time-frequency spectral analysis of heart rate variability during induction of general anaesthesia. British journal of anaesthesia, 79(6):754 -- 758, 1997. [47] Xiao J and Flandrin P. Multitaper time-frequency reassignment for nonstationary spectrum estimation and chirp enhancement. Signal Processing, IEEE Transactions on, 55(6):2851 -- 2860, 2007. [48] Mazzeo AT, La Monaca E, Di Leo R, Vita G, and Santamaria LB. Heart rate vari- ability: a diagnostic and prognostic tool in anesthesia and intensive care. Acta Anaes- thesiologica Scandinavica, 55(7):797 -- 811, 2011. [49] Thakur G, Brevdo E, Fuckar NS, and Wu HT. The synchrosqueezing algorithm for time-varying spectral analysis: robustness properties and new paleoclimate applica- tions. Signal Processing, 93(5):1079 -- 1094, 2013. [50] Flandrin P. Time-frequency/time-scale analysis, volume 10. Academic Press, 1998. [51] Daubechies I et al. Ten lectures on wavelets, volume 61. Philadelphia: SIAM, 1992. [52] Chui CK. Wavelets: A Mathematical Tool for Signal Analysis, volume 1. SIAM, 1997. [53] Chui CK. An introduction to wavelets, volume 1. Academic press, 1992. [54] Kodera K, Gendrin R, and Villedary C. Analysis of time-varying signals with small bt values. Acoustics, Speech and Signal Processing, IEEE Transactions on, 26(1):64 -- 76, 1978. [55] Auger F and Flandrin P. Improving the readability of time-frequency and time-scale representations by the reassignment method. Signal Processing, IEEE Transactions on, 43(5):1068 -- 1089, 1995. Bibliography 71 [56] Daubechies I and Maes S. A nonlinear squeezing of the continuous wavelet transform based on auditory nerve models. Boca Raton, FL, USA: CRC Press, 1996. [57] Wu HT, Flandrin P, and Daubechies I. One or Two Frequencies? The Synchrosqueez- ing Answers. Advances in Adaptive Data Analysis, 3:29 -- 39, 2011. [58] Thakur G and Wu HT. Synchrosqueezing-based recovery of instantaneous frequency from nonuniform samples. SIAM Journal on Mathematical Analysis, 43(5):2078 -- 2095, 2011. [59] Wu HT. Instantaneous frequency and wave shape functions (i). Applied and Compu- tational Harmonic Analysis, 35(2):181 -- 199, 2013. [60] Wu HT, Chan YH, Lin YT, and Yeh YH. Using synchrosqueezing transform to discover breathing dynamics from ecg signals. Applied and Computational Harmonic Analysis, 36(2):354 -- 359, 2014. [61] Auger F, Flandrin P, Lin YT, McLaughlin S, Meignen S, Oberlin T, and Wu HT. A coherent overview of time-frequency reassignment and synchrosqueezing. IEEE Signal Processing Magazine, 30(6):32 -- 41, 2013. [62] Steven MK. Modern Spectral Estimation: Theory and Application. Prentice-Hall, N.J., 1988. [63] Thomson DJ. Spectrum estimation and harmonic analysis. Proceedings of the IEEE, 70(9):1055 -- 1096, 1982. [64] Bayram M and Baraniuk RG. Multiple window time-varying spectrum estimation. Nonlinear and Nonstationary signal processing, pages 292 -- 316, 2000. [65] Chui CK and Chen G. Kalman filtering: with real-time applications. Springer, N.Y., 2009. [66] Sheiner LB, Stanski DR, Vozeh S, Miller RD, and Ham J. Simultaneous modeling of pharmacokinetics and pharmacodynamics: application to d-tubocurarine. Clinical pharmacology and therapeutics, 25(3):358 -- 371, 1979. [67] Smith WD, Dutton RC, and Smith NT. Measuring the performance of anesthetic depth indicators. Anesthesiology, 84(1):38 -- 51, 1996. [68] Wolfe R and Hanley J. If we're so different, why do we keep overlapping? when 1 plus 1 doesn't make 2. Canadian Medical Association Journal, 166(1):65 -- 66, 2002. [69] Lin YT, Wu HT, Tsao J, Yien HW, and Hseu SS. Time-varying spectral analysis re- vealing differential effects of sevoflurane anaesthesia: non-rhythmic-to-rhythmic ratio. Acta Anaesthesiologica Scandinavica, 58(2):157 -- 167, 2014. Bibliography 72 [70] Blues CM and Pomfrett CJ. Respiratory sinus arrhythmia and clinical signs of anaes- thesia in children. British journal of anaesthesia, 81(3):333 -- 337, 1998. [71] Galletly DC, Westenberg AM, Robinson BJ, and Corfiatis T. Effect of halothane, isoflurane and fentanyl on spectral components of heart rate variability. British journal of anaesthesia, 72(2):177 -- 180, 1994. [72] Sleigh JW and Donovan J. Comparison of bispectral index, 95% spectral edge fre- quency and approximate entropy of the eeg, with changes in heart rate variability during induction of general anaesthesia. British journal of anaesthesia, 82(5):666 -- 671, 1999. [73] Luginbuhl M, Ypparila-Wolters H, Rufenacht M, Petersen-Felix S, and Korhonen I. Heart rate variability does not discriminate between different levels of haemodynamic responsiveness during surgical anaesthesia†. British journal of anaesthesia, 98(6): 728 -- 736, 2007. [74] Tanaka M and Kunimatsu J. Contribution of the central thalamus to the generation of volitional saccades. European Journal of Neuroscience, 33(11):2046 -- 2057, 2011. [75] Bimer J and Bellville JW. Arousal reactions during anesthesia in man. Anesthesiology, 47(5):449 -- 454, 1977. [76] Gunnarsson F and Gustafsson F. Frequency analysis using non-uniform sampling with application to active queue management. In Acoustics, Speech, and Signal Processing, 2004. Proceedings.(ICASSP'04). IEEE International Conference on, volume 2, pages ii -- 581. IEEE, 2004. [77] Vio R, Strohmer T, and Wamsteker W. On the reconstruction of irregularly sampled time series. Publications of the Astronomical Society of the Pacific, 112(767):74 -- 90, 2000. [78] de Boor C. A practical guide to splines. Springer-Verlag, 1978. [79] de Boor C and Fix GJ. Spline approximation by quasi-interpolants. Journal of Approximation Theory, 8:96 -- 110, 1973. [80] Lyche T and Schumaker LL. Local spline approximation methods. Journal of Ap- proximation Theory, 15(4):294 -- 325, 1975. [81] Schumaker LL. Spline functions: Basic theory. Wiley-Interscience, 1981. [82] de De Villiers J. Mathematics of Approximation. Atlantis Press, 2012. [83] Chen G, Chui CK, and Lai M. Construction of real-time spline quasi-interpolation schemes. Approx. Th. and its Appl., 4:61 -- 75, 1988. Bibliography 73 [84] Chui CK and Diamond H. A general framework for local interpolation. Numerische Mathematik, 58(1):569 -- 581, 1990.
1911.04585
1
1911
2019-11-11T22:21:22
Modulation of oculomotor control during reading of mirrored and inverted texts
[ "q-bio.NC" ]
The interplay between cognitive and oculomotor processes during reading can be explored when the spatial layout of text deviates from the typical display. In this study, we investigate various eye-movement measures during reading of text with experimentally manipulated layout (word-wise and letter-wise mirrored-reversed text as well as inverted and scrambled text). While typical findings (e.g., longer mean fixation times, shorter mean saccades lengths) in reading manipulated texts compared to normal texts were reported in earlier work, little is known about changes of oculomotor targeting observed in within-word landing positions under the above text layouts. Here we carry out precise analyses of landing positions and find substantial changes in the so-called launch-site effect in addition to the expected overall slow-down of reading performance. Specifically, during reading of our manipulated text conditions with reversed letter order (against overall reading direction), we find a reduced launch-site effect, while in all other manipulated text conditions, we observe an increased launch-site effect. Our results clearly indicate that the oculomotor system is highly adaptive when confronted with unusual reading conditions.
q-bio.NC
q-bio
Modulation of oculomotor control during reading of mirrored and inverted texts Johan Chandra, André Krügel, Ralf Engbert Deparment of University of Potsdam, Potsdam, Germany November 13, 2019 Abstract The interplay between cognitive and oculomotor processes during reading can be explored when the spatial layout of text deviates from the typical display. In this study, we investigate various eye-movement measures during reading of text with experimentally manipulated lay- out (word-wise and letter-wise mirrored-reversed text as well as in- verted and scrambled text). While typical findings (e.g., longer mean fixation times, shorter mean saccades lengths) in reading manipulated texts compared to normal texts were reported in earlier work, little is known about changes of oculomotor targeting observed in within- word landing positions under the above text layouts. Here we carry out precise analyses of landing positions and find substantial changes in the so-called launch-site effect in addition to the expected overall slow-down of reading performance. Specifically, during reading of our manipulated text conditions with reversed letter order (against overall reading direction), we find a reduced launch-site effect, while in all other manipulated text conditions, we observe an increased launch- site effect. Our results clearly indicate that the oculomotor system is highly adaptive when confronted with unusual reading conditions. 1 Introduction Visual acuity is greatest at the center of visual field (the fovea) and declines sharply on the periphery, which limits the information extraction process of visual input from the environment. To compensate the limitation, the eyes generate short and rapid movements, saccades, to shift the fovea to the regions of interests for high-acuity information processing [11]. Similar principle is observed during processing of writtten text or reading. During reading, the eyes typically move forward about 6-7 character spaces dur- ing saccades and fixate on a word for about 200-250 ms to support word processing. The control of (saccadic) eye movements requires the coordina- tion of several fundamental cognitive subsystems such as word recognition, attention [8], and oculomotor control [45]. While the cognitive system is responsible for selecting which word to be fixated next, it is the oculomotor system that is responsible for shifting the fovea to the regions of interests for high-acuity information processing [11]. Thus, our reading ability depends on the oculomotor performance, whose properties are reflected most clearly in the statistics of within-word fixations, typically the eyes' landing position on words after saccades. Unlike the well-documented effects of cognitive modulation on temporal aspects (e.g. fixation duration) of eye movement measures [e.g. 14, 40], small effects of cognitive modulation on spatial aspects (e.g. within-word landing position) of eye movement measures were reported. For example, ortho- graphic familiarity and regularity influence landing positions [12, 55, 56, 57]. Furthermore, corpus analyses showed a significant effect of word frequency on mean fixation position: Saccades landed further into the (3- to 6-letter) tar- get word, when it was a high-frequency word as compared to a low-frequency word [29]. Lavigne [24] reported a shift of initial fixation location toward the end of high predictable words, but only when the words were seen more frequently and for saccades lauched near the word beginning. On the other hand, Rayner [41] reported that word predictability had little influence on initial landing position on word and suggested that landing position effects in reading were primarily modulated by low-level processing. Finally, data from z-string scanning (where all letters were replaced by the letter "z" or "Z") indicated that within-word landing position distributions are very sta- ble and do not critically depend on meaningful content ([31]; see also [27]). Moreover, most effects of higher-level processing on mean fixation position are small (< 0.5 character spaces). Interestingly, the within-word landing position were reported to influence the "higher-level" processes. In several studies, it was reported that (isolated) word recognition time was at the minimum when the eyes land at the center of the word compared to when the eyes land on the word's periphery, termed 2 the optimal viewing position [33, 34]. Similar but weaker effect was observed on refixation probability in reading: readers were less likely to refixate the words if fixation land near the word center [53]. Oppositely, mean single fixation duration is the greatest when the eyes land at word center, termed the inverted optimal viewing position in Vitu et al. [52]. To explain the effect, Nuthmann and colleagues [30] argued that fixations landed on the word egdes were typically not intended to land on the target (mislocated fixations), reflected on the short durations. The observed evidences in reading researches are the result of the inte- gration of cognitive processes in word selection and oculomotor processes in shifting the eyes to the area of interest. To isolate the underlying process affected the observation, one should manipulate factors associated to one process while controling the remaining factors. In fact, Kolers and Perkins [15] used geometric rotations, reflections and other transformations of text as the physical variation to study the recognizability of the texts and the influence that practice in reading one type of transformation applied on the recognition of others. They found that the transformations being tested var- ied in difficulty and transferability. Likewise, Kowler and Anton [19] applied similar types of transformation to test the effects on global saccade lengths and fixation durations. By observing eye movement patterns of two partic- ipants, they reported that the directional pattern of saccades had relatively modest effects on reading speed under the instruction to read accurately. They argued that the reading time was affected by longer time needed to generate short saccades observed in reading difficult texts. In a separate test, they found negative relationship between saccade length and saccade latency: short saccades (less than 30') have longer latency than long saccades. Additionally, as a response to the Internet myth, Rayner et al. [42] tested different types of word transposition (internal, beginning, and end of word) on reading time and reported that although participants were able to read the text, reading time was slower for some transposition types, especially when the word beginning was transposed. Hence they concluded the impor- tance of word beginning (see also [54] for similar conclusion). Following the above approach of presenting texts in unfamiliar representation, we designed a study with four different experimental conditions and a control condition to systematically investigate the possible modulations of oculomotor pro- cesses in response to variations of the spatial layout of texts. Furthermore, the current study employs various eye movement measures to describe the oculomotor performances during reading. 3 In current study, letter positions and word representations were exper- imentally manipulated in the following ways: We used texts composed of mirror-inverted letters (mL), mirror-inverted words (mW), inverted words (iW), where regular letters are printed in reverse order, and scrambled letters (sL) (see Figure 1a for an example). In mirrored-words (mW) and mirrored- letter (mL) conditions, either the complete word or the constituting letters were mirror-inverted with respect to the vertical axis. In contrast, no mirror- ing was involved in the construction of inverted-word (iW) and scrambled- word (sL) text conditions. In inverted words (iW) condition, letter represen- tation was normal, but the position was inverted in the iW condition to mimic the letter position in the mW condition. In scrambled letters (sL) condition, the positions of the first and the last letter of a word were maintained, while the letter positions in between were randomized (i.e., there was no change in words with length less than 4 letters). The condition mL, mW and iW are equivalent to the condition NNV, NRV and NRN conditions in [19] while the condition sL is equivalent to the internal transposition manipulation in [42]. The following sections describe the robust finding on within-word landing position distribution during reading and the proposed models to explain the observed phenomena. Furthermore, we will discuss about current reading models and their predictions, particularly on saccade generation and "where" the eyes land, in relation to the manipulations in current study. Within-word landing positions Regarding where the eyes land during reading, a robust finding is that within- word landing position approximately follows a Gaussian density function, with a pronounced peak, typically located halfway between word beginning and word center, but with a surprisingly large variance [39, 28, 32]. The landmark study by McConkie et al. [28] identified two independent oculo- motor error components in reading, which we will denote as the range-error model throughout this article: (i) The random placement error is assumed to reflect perceptuo-oculomotor inaccuracy in the execution of saccades, which can be approximated by a Gaussian distribution. (ii) The saccadic range error represents a systematic, launch-site contingent shift of mean landing positions and is typically explained as a general response bias of the human motor system [36]. Specifically, McConkie and colleagues [28] found that during reading, the within-word landing positions varied systematically with the launch-site dis- 4 (a) (b) Figure 1: Experimental sentence stimuli and hypotheses on saccade lengths. (a) In the control condition, normal German text (N) is presented; for mir- rored letters (mL) and scrambled words (sL), we expect shorter saccade lengths, on average. For mirrored words (mW) and inverted words (iW), word beginnings and word ends are exchanged, so that we expect longer mean saccade lengths and more regressions due to more frequent re-readings of the same string. (b) The distribution of saccade lengths (solid lines) and the expected changes due to the experimental manipulations. 5 JedeSprachederWeltbesitzteineGrammatik.(N)(mL)(mW)(iW)(sL)JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.WeltbesitzteineGrammatik.Legende:!!!!!!!!!Vorwärtssakkade,!!!!!!!!!Refixa1on,!!!!!!!!!Skipping,!!!!!!!!!!!Regression!!(N)!!!(G)!!!(bG)!JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.JedeSprachederWeltbesitzteineGrammatik.edeJehcarpSredtleWtztisebeniekitammarG.JdeeScrahpederWletbsizettenieGmartimak.-10-505101520Saccade length [char]Relative frequencyForward saccadesRegressionsShorter saccades (all)More Regressions (mW, iW) tance, i.e., the distance between fixation location (before the saccade) and the beginning of the target word. The stable observation in reading is that each letter increment of the saccade launch-site generates a shift of mean landing-site with a magnitude of about half a character space to the left, the launch-site effect, which is independent of the target word length. If the dis- tances between launch-sites and landing sites are measured relative to word centers, then the within-word mean landing position can be described by a linear landing-position function [see also 38] of the form ∆P V P = λ · (L0 − L) , (1) where ∆P V P denotes the average shift of the within-word mean landing po- sition from the word center and L is the distance between launch site and the center of the target word. While a negative value of ∆P V P indicates that the within-word mean landing position shifts to the left of the word center, a positive value indicates a rightward shift. The parameter L0 represents the center-based launch-site distance, where saccades land precisely on the word center, on average. The strength of the launch-site effect is represented by the slope parameter λ. An estimated slope of λ ≈ 0.5 was observed in readers of English [28] and German [30] texts. While the range-error model was generally a successful first description of the eye-movement data, experimental studies demonstrated effects that could not be explained within this model. First, the presence of additional stimuli could influence the saccadic landing positions [3, 4, 10, 49, 50, 51]. Second, Krügel and Engbert [20] demonstrated that saccade type (i.e. word skipping) could influence the saccade landing positions during reading [see also 22, 37]. These findings challenged the range-error model in explaining observed saccadic landing positions during reading. It is important to note, however, that the range-error model is purely descriptive, since it does not include more fundamental computational prin- ciples for oculomotor control. Furthermore, the slope parameter of λ repre- sents quantification of the strength of the launch-site effect, without direct inferences on what processes underlying the observed phenomenon. As a consequence, it is not surprising that integrating new experimental evidence in the range-error model is difficult. In the next section, we discuss a process- oriented Bayesian model of within-word fixation position that was developed over the last 10 years. 6 A Bayesian model of oculomotor control in reading The framework of Bayesian decision theory has been proposed as a principled approach to the optimal control of human behavior in the context of integra- tion of sensorimotor and cognitive processes [16, 17, 18]. Since saccadic eye movements require both sensorimotor (i.e., moving the eyes to foveate words) and cognitive processes, Engbert and Krügel [5] proposed that eye movement control during reading could be explained using Bayesian estimation. Accord- ing to Bayes rule [e.g., 58], the optimal estimate of a target position x, given a sensory observation at x0, can be calculated as the conditional probability (posterior) π(xx0) ∼ q(x0x)p(x) , (2) where p(x) is the previously learned prior distribution of the target, inde- pendent of current sensory input, and the conditional probability q(x0x) is the sensory likelihood of the observation at position x0 given a target at x. The relation in Eq. (2) determines the dependence of the posterior proba- bility from x. The missing constant of proportionality can be obtained by normalization. Engbert and Krügel [20] proposed that the dependence of landing po- sitions within a word on the saccadic launch site is a special case of the Bayesian principles, Eq. (2), for saccade planning in reading. In a mathe- matical model, the likelihood q(x0x) was modeled as an unbiased, normally distributed probability density centered at the intended target word and with 0, which represents the degree of sensory uncertainty. Assuming variance σ2 that the prior distribution is also normally distributed, the product of the two Gaussian densitiy functions results in another normal density function, the posterior probability density π(xx0). The posterior probability provides a natural explanation of the launch-site effect, since the position of its maxi- mum falls between the maximum of the prior distribution and the maximum of the sensory likelihood. As a result, the posterior reproduces the systematic tendencies of saccades (i) to overshoot the center of close target words and (ii) to undershoot the center of distant target words [5]. The shift of the mean of the posterior µP from the observation x0 can be calculated as ∆Bayes = µP − x0 = σ2 0 0 + σ2 σ2 T (µT − x0) , (3) where σ2 T is the variance of the prior probability of center-based launch- site distances. Comparing the equation for the launch-site effect, Eq. (1), 7 with the predicted effect in the Bayesian theory, Eq. (3), thus assuming ∆P V P ≡ ∆Bayes,we obtain . (4) λ = σ2 0 σ2 0 + σ2 T Recently, Krügel and Engbert [21] introduced an advanced model, which includes an explicit model for the computation of the word center from sen- sory estimates of word boundaries. Therefore, Bayesian models provide a robust theoretical framework to explain where the eyes move during reading. One advantage to model within-word landing position distribution using an explicit Bayesian model is that the interpretability of the results. The slope parameter λ estimated from Bayesian model represents weighting of optimal behavior during reading: maintaining constant saccade length while target- ing word center. For extreme case where λ → 0, it can be interpreted that the optimal oculomotor control in reading puts more weight on the precision in landing on target precision, hence minimizing range error. On the other hand, the value of λ → 1 means that optimal reading behavior rely on main- taining constant saccade length, reducing the importance of target location. With the typical empirical value of λ ≈ 0.5, we obtain the important result that the sensory variance of the target location is approximately the same as the variance of the prior distribution, i.e., σ2 T , in optimal reading behavior. 0 ≈ σ2 Hypothesis and predictions from various reading models Reading models were typically developed to help understanding the complex processes underlying reading processes. Despite the fact that eye movement control during reading requires both cognition and oculomotor systems, as mentioned above, empirical studies found that cognition had small effects on within-word landing positions. Interestingly, adding additional visual cue [? ] or changing sentence presentation, i.e. texts are read from top to bot- tom [13] did not notably change the landing position distributions on words. Consequently, most mathematical models of saccade generation during read- ing assume that oculomotor control is dissociated from cognitive processes. Cognitive-based reading models (e.g., E-Z Reader: [44, 43, 46]; SWIFT: [6, 8, 47]) assume that cognitive processes related to language processing are responsible for eye movements without distinguished effects on oculomotor within-word targeting process. Moreover, most cognitive models based their 8 implementation of saccadic errors on the range-error model ([28]) with rel- atively fixed values. Since the manipulation types tested in current study maintain the spatial information such as word length, cognitive-based mod- els will not predict substantial changes in within-word landing positions since they asumme that oculomotor process is mainly affected by "low level" in- formation. However, since text manipulations will increase processing loads, hence affecting saccade generating time, these models will generate more refixation saccades but less skipping saccades. Most relevant to current study is Mr.Chips [26, 25], an ideal-observer model that combine visual, lexical, and oculomotor information optimally to read simple texts in the minimum number of saccades. The model op- erates according to an entropy-minimization principle, generating saccades that minimize uncertainty about the current word or saccades that move the visual span furtherst to the right. Note that for Mr.Chips, a word is said to be fixated if the central slot of the visual span (a linear array of character slots with each slot can be either high or low resolution) falls on one of the letters of the word and this central slot does not have preferred status. Mr. Chips' skipping rate and global landing position distribution were similar to human data. Mean saccade length decreases as the lexion size increases. Further- more, it generated less refixations and reduced launch-site effect (λ = 0.21) in reading normal text. Given that word identification played a key role for lexical processing in the model, we speculate that if the words were written from right to left (e.g. in mW and iW conditions), Mr. Chips should gen- erate more saccades that land on the second half of a word to capture more information about word identity, assuming that the first half of the word were identified beforehand and letter mirroring and inversion do not affect it's lexical access process. Our hypothesis for eye-movement measures on reading unfamiliar typog- raphy are derived from the predicted increase in perceptual difficulty and additional oculomotor demand. Longer fixation durations and shorter sac- cade amplitudes can be expected for more difficult texts in all four conditions. Increased perceptual difficulty should also result in less skipping cases, but more refixations; both of these predictions are compatible with a reduced average saccade length (Fig. 1b). On the level of within-word landing postions, our hypotheses are more specific for the different experimental conditions. Reading words with mirror- inverted letters (mL) or with scrambled letters (sL) will produce within- word landing positions similar to normal reading, since information on letter 9 positions did not change (mL) or did not deviate systematically from nor- mal reading (sL). However, due to inversion of letter positions in reading mirrored-words (mW) and inverted words (iW), we expect readers to shift their eyes further to the right of the word string in the initial saccade and to generate a regressive refixation after that inital saccade (see Fig. 1a). Our study employs a learning paradigm, where participants are required to train reading text in one of the four experimental conditions. The mo- tivation for the learning paradigm is to check the stability of the resulting eye-movement patterns and to exclude the possibility that the results are a short-lived effect due to first exposure to an unfamiliar layout. In general, we expect that the training will lead to improved performance on the level of global reading measures (average fixation times and mean saccade lengths). Since the scrambled letters (sL) condition represents the least systematic variation of the text layout, we expect that the learning effects are smaller than in the other three experimental conditions. On the level of within-word landing positions, we expect that stable shift of the initial landing positions towards the end of the word strings will be established during learning in the conditions with mirrored words (mW) and inverted words (iW), since the word beginnings are at the end of the manipulated string in these two conditions. For the other two conditions (mL, sL), we expect that within-word landing postions are very similiar to normal reading after training. 2 Results Reading text with manipulated layout due to mirrored-reversed and inverted letter arrangement produced changes in behavioral measures on a global level (e.g., mean fixation duration, average saccade length) and on the oculomtor level (i.e., within-word fixation locations/saccadic landing positions). While the primary goal of our study was to investigate the possible modulations of oculomotor processes in response to variations of the spatial layout of texts, we start with the presentation of results on global summary statistics to evaluate the overall effects on reading performance as characterized by eye-movement measures. 10 Global summary statistics This section highlights the significant results for each manipulation across training session in comparison with normal reading. If not otherwise men- tioned, results refer to comparisons between the first experimental session (in one of the four conditions mW, iW, mL, or sL) and the control condition (normal reading). First of all, global summary statistics from reading normal texts are repli- cated. In control condition, most of the first-pass fixations move forward to the next neighbouring words (forward saccades 41%) or skip (skipping sac- cades 26%). About 22% of the fixations move within a word (refixation saccades) and 12% of them move backward (regression saccades). On av- erage, readers move their eyes 7.82 character spaces forward, 4.23 character spaces backward and fixated on words for 245 ms during reading normal text. Compared to control condition, percentage of forward saccades in experimen- tal condition did not show strong deviation (31% - 44%). The percentage of skipping saccades was reduced to less than 5% in conditions where texts were written against the reading direction (e.g. in mW and iW conditions) and between 10-15% in other conditions. Interestingly, half of the first-pass fixations observed in conditions where texts were written against reading direction came from refixation saccades, while only a third of first-pass fixa- tions were generated from refixation saccades. Less than 12% of fixations in reading manipulated text were moving backwards (e.g. regression saccades) with the lowest observed in iW condition. When reading the manipulated texts for the first time, readers fixated on words, on average, about 53-137 ms longer and generated shorter saccades (5-6 character spaces for progressive saccades and 3 character spaces) than those observed in reading normal text. Note that mean fixation duration in sL condition is about 50 ms longer than those reported in the study by White and colleagues [54]. Global summary statistics from all sessions are presented in Table A1. On a qualitative level, we compare the resulting saccade length distribu- tions (Fig. 2) with our hypotheses (Fig. 1b). As expected, the distributions of foward saccades length during reading manipulated text are generally shifted to the left of those from control condition, indicating a qualitatively shorter saccade length generated. Futhermore, more regressive saccades were ob- served only in reading mirrored-word (mW) and inverted-word (iW) condi- tions. 11 Figure 2: Distribution of saccade amplitudes across all sessions. Saccades observed in the control condition (normal reading) are marked with red color. Dark blue lines represent data from first session of reading manipulated text. Data from sessions after the two trainings are marked with lighter blue hues. 12 Inverted Words (iW)Scrambled Letters (sL)Mirrored Words (mW)Mirrored Letters (mL)−20−1001020−20−10010200.000.050.100.000.050.10Saccade Length [letters]Relative FrequencySessionControlSession 1Session 2Session 3 Single fixation duration and refixation probability Results from statistical modeling using linear mixed-effects models for single fixation duration and refixation probability are summarized in Tables A2-A4. For specification of linear mixed-effects models, see section 4. No significant word length effect was observed in single fixation duration measure for control condition, which in line with the finding reported in study conducted by Kliegl et al. [14]. Compared to control condition, word-based single fixation durations observed in experimental conditions are significantly longer (all t value > 2). For example, on medium words (5-7 characters), estimated mean single fixation durations in the first experimental sessions were 48-104 ms longer than in control condition (estimated mean: 238 ms). In line with the study by Kowler and Anton [19], texts written against reading direction were more difficult to process. This processing difficulties seemed to be the property of the corresponding type of text manipulation and were more obvious after participants have learned how to read the texts. When readers fixated on medium words only once in the last experimental sessions, the fixation durations in mW (estimated mean: 309 ms) and and sL (estimated mean: 326 ms) conditions were shorter compared to those observed in mW and iW conditions with estimated mean of 570 ms and 455 ms respectively. Similar to single fixation measures, there was no significant word length effect observed in first of multiple fixation duration measures for control con- dition. The mean of first fixation duration in sL condition did differ signifi- cantly across all sessions compared to those observed in control condition. In mL condition, the duration of first fixation measures was significantly longer when readers read the manipulated text for the first time, but did not differ significantly after trainings. Interestingly, the mean of first fixation duration was significantly longer in mW and iW conditions compared to control condi- tion. For example, on medium words, estimated mean durations of the first fixation in both conditions were about 90 ms longer than those estimated for control condition. The difference remained significant even after trainings (estimated means session 3: mW = 465.32; iW = 414.38). On refixation probability measure, we replicated the word length effect in both control and experimental conditions (all p < 0.01). Long words were more likely to be refixated than short words. However, when the texts were manipulated, readers were more likely to refixated the words. For example, the estimated refixation probability on medium words in control condition was 0.16. However, readers in experimental conditions were twice more likely 13 to fixate the words with the same lengths (estimated RFP: mL = 0.42; sL = 0.65; mW = 0.44; iW = 0.56). Interestingly, after training to read texts with words written against the reading direction (e.g. mW and iW conditions), readers were almost always fixated on medium and long words (estimated RFP above 0.87 on the third session). Do these manipulation types have something in common or do they gen- erate different effect on fixation duration and probability measures presented above? The results from separate models which estimated the different effect size of manipulation types (see Table A5) confirmed that some manipulation type generated greater effect than the others. All of the three models showed similar trends. The effect in mL condition was significantly different from control condition, but adding the effect from sL conditon to the mean of the two conditions (control and mL conditions) did not yield a significant gain on effect size. However, adding the effect from mW condition and iW conditions yielded significant gains on the effect size. The results from various measures showed that of all manipulation types tested in the current study, the iW condition generated the largest effect. The effects of word lengths on duration of single fixation and first of multi- ple fixations, as well as on skipping and refixation probabilites are visualized in Fig. 3. Landing positions distributions and launch-site effect A first glance at the resulting distributions for within-word landing positions indicates differences between normal reading and manipulated texts (Ap- pendix A3-A4). Except for short words (i.e., word length up to 4 chars), we found increased leftward shifts across all experimental conditions, where with mirrored words (mW) and inverted words (iW) showed stronger shifts com- pared to the other two manipulations (mL, sL). Since shorter saccades were generated in experimental conditions, we analyses only saccades launched up to 5 character away from the word beginning, resulting in more overshoots. Undershoots were typically observed on longer saccades launched furhter than 7 characters away and more prominent on long words [28, 30, 20]. Based on Bayesian fits of the distribution (see Methods), we obtained mean landing positions that were used for further analyses. In Figure 4a, center-based mean landing sites are plotted against center- based launch sites for forward saccades and different lengths of the target words. The slope parameters from landing position functions are plotted for 14 Figure 3: Upper row. Mean fixation duration of single fixation (left) and first of multiple fixations as the function of word length classes. Lower row. Skipping (left) and refixation probabilities as the function of word length classes. Word with length less than 3 characters are grouped in word length class 3; long words (> 8characters) are grouped in word length class 8. Error bars represent the standard error of the means. Red line and dots represent data from reading normal text. The dark blue color represents data from the first experimental session. The lighter blue hues indicate data from the last two experimental sessions. 15 llllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllInverted Words (iW)Scrambled Letters (sL)Mirrored Words (mW)Mirrored Letters (mL)345678345678200300400500200300400500Word length classMean Fixation Duration [ms]Single FixationllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllInverted Words (iW)Scrambled Letters (sL)Mirrored Words (mW)Mirrored Letters (mL)345678345678200300400500200300400500Word length classMean Fixation Duration [ms]First of Multiple FixationsllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllInverted Words (iW)Scrambled Letters (sL)Mirrored Words (mW)Mirrored Letters (mL)3456783456780.000.250.500.751.000.000.250.500.751.00Word length classSkipping probabilitySkipping Probability on Word LengthllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllllInverted Words (iW)Scrambled Letters (sL)Mirrored Words (mW)Mirrored Letters (mL)3456783456780.000.250.500.751.000.000.250.500.751.00Word length classRefixation probabilityRefixation Probability on Word LengthSessions:llllControlSession 1Session 2Session 3 (a) (b) Figure 4: Analysis of the launch-site effect. (a) Landing position function: Mean center-based landing position as a function of center-based launch- site distance for all experimental conditions. Red line and dots represent data from normal reading session. Data from the first experimental sessions are presented in dark blue color. The lighter blue hues represent the last two experimental sessions. (b) The estimated slope parameter λ (and the 95 % confidence interval) for all experimental conditions as a function of the experimental session. The red horizontal dashed line represents the estimated slope for reading normal text (baseline). Results for reading mirrored words (mW) and inverted words (iW) are presented with magenta and cyan lines, resp. Green and yellow lines represent estimations for reading texts with mirrored letters (mL) and scrambled letters (sL). 16 Inverted Words (iW)Scrambled Letters (sL)Mirrored Words (mW)Mirrored Letters (mL)−9−8−7−6−5−4−3−9−8−7−6−5−4−3−10123−10123Launch−site distance to word centerCenter−based mean landing positionSession:ControlSession 1Session 2Session 3Landing Position FunctionControlControlControlControlControlControlControlControlControlControlControlControlControlControlControlControl0.20.30.40.50.6Session 1Session 2Session 3SessionSlopeCondition:mWmLiWsLEffect on Slope of Landing Position Function all sessions in Figure 4b. The estimated slope of (λ0 = 0.37) was observed in reading normal text. Compared to the value from control condition, both conditions iW (λ1 = 0.31,λ2 = 0.25,λ3 = 0.23) and mW (λ1 = 0.24, λ2 = 0.26, λ3 = 0.29) generated shallow slopes, which indicates a tendency to reduced oculomotor control. In contrast, conditions mL (λ1 = 0.44, λ2 = 0.38, λ = 0.37) and sL (λ1 = 0.52, λ2 = 0.48, λ = 0.41) generated greater slope values (λ) compared to reading normal text, indicating a tendency to increased oculomotor control. Interestingly, except in the mW condition, the λ values approach the value of normal reading after trainings. 0) and standard deviation of prior distribution (σ2 In the Bayesian model, the slope parameter λ represents the strength of oculomotor control and the relation between observational error of the target T ). In extreme location (σ2 0 → 0, the eyes always land on the cases, a value of slope near to 0, where σ2 target (i.e., the word center) regardless of where the saccade started. In the other extreme case, the a slope value of λ → 1 indicates the absence of a target selection process in saccade planning, so that the eyes generate random constant saccade lengths (from a uniform distribution). Therefore, our data show that saccades from reading text composed of words with reversed letter sequences (i.e., mW and iW) tend to land precisely on the target location (word center) on average, while reading manipulated text with normal letter order (i.e., mL and sL) tend to generate saccades similar length. Comparing the results with our hypotheses, there is no dramatic effect of a shift of the mean landing position towards the word ends. Therefore, we ran a post-hoc analyses for single and two-fixation cases in the next section. Landing positions for single vs. two-fixation cases Single fixation cases. Only cases where exactly one fixation land on a word were considered. Relative frequency of fixation landing position were cal- culated based on word length. To obtain estimates for the mean µSF and standard deviation σSF for the landing position distribution of each word length, a grid search method (in steps of 0.01) with a minimum-χ2 criterion was applied. Analysis of variance was conducted to statistically compare the manipulation effects on mean (µSF ) and standard deviation (σSF ) of landing position distributions. No significant difference was observed for the mean (µSF ) of the landing position distribution of single fixation cases between control and experimen- tal conditions (F [1, 30] = 2.364, p = 0.135). The main effect on standard 17 deviation (σSF ) of within-word landing position distribution of single fixation cases yielded an F -ratio of F [1, 30] = 32.7,p < 0.000, indicating a significant difference between reading normal text (M=2.95,SD=0.84) and manipulated text (M=1.51,SD=0.55). When words were fixated only once, readers' mean landing position did not change across the manipulation types. However, the precision of landing on selected target increased as the text display deviated from normal presentation. Among our key results is that the two conditions where the word begin- nings were located at the end of the manipulated word strings (i.e., conditions mW and iW) produced only a slight shift of the average landing position. This finding turned out to be robust and remained observable even after trainings. Figure 5 presents data for single fixations. The finding did not support our hypothesis that readers targeted the second half of the word strings during reading mirrored-word (mW) and inverted words (iW). On a qualitative level, there is little adaptivity of the oculomotor system. Since in our hypothesis, an initial saccade towards the second half of the word would require an additional refixation in the first half of the word, we ran a post-hoc analysis of all two-fixation cases. The corresponding plot in Figure 6 presents data for the initial landing position (first saccade into the word) from all cases, where the word was fixated exactly two times. In contrast to typical OVP effect on refixation probability, plotting the initial of two fixations give us a better understanding of word targeting in saccade planning. Two-fixation cases. For this analysis, we considered cases when exactly two fixations on a word were observed. The landing position distribution of the initial fixation was fitted to a quadratic polynomial, i.e., y = A + B · L(x − C)2 , (5) where x denotes the initial landing position and y is the relative frequency of the fixation. The parameter A represents the actual relative frequency of the initial landing position. The parameter B is the slope of the parabolic curve, representing the within-word maximum or minimum relative frequency of the landing position. The parameter C reflects within-word position where the relative frequency was at the minimum or maximum, depending on the value of parameter B. In general, the value of the parameter B is assumed to be positive, resulting in a distribution of landing position qualitatively similar to the optimal viewing position (OVP) curve but with different interpretation: 18 Figure 5: Within-word landing position distributions for single-fixation cases. Columns relate to different word lenghts. Rows refer to experimental manipulations. Different colors indicate control condition (red) and experi- mental conditions (blue hues). 19 4−letterwords5−letterwords6−letterwords7−letterwordsMirrored Words (mW)Mirrored Letters (mL)Inverted Words (iW)Scrambled Letters (sL)02460246024602460.00.10.20.30.40.00.10.20.30.40.00.10.20.30.40.00.10.20.30.4Landing PositionRelative Frequency Session: ControlSession 1Session 2Session 3Single Fixation Cases when a word is fixated exactly twice, the initial fixations tend to on word edges more often than on the word center. The estimation of the three parameters representing the characteristics of the landing position of the first of two fixation cases was conducted based on a maximum likelihood method using the bbmle package [2] for R studio. The estimated parameters are summarized in Table A6-A10. In two-fixation cases under normal reading conditions, readers tend to ini- tially land their eyes close to the word beginning with a subsequent secondary fixation further into the word (most refixations are forward directed). Sim- ilar patterns were observed in the condition where manipulated texts were written in typical direction from right to left (mL and sL). However, what is most remarkable for our study, in the two conditions (mW and iW) with reversed letter sequences, the initial landing positions deviated from those of normal reading (for detailed estimated parameters, see Table A7 and A9). When a word was fixated more than once, the first saccade landed mostly behind the location of the word center in the second half of the word strings; thus, there is clear top-down adaptivity in the two conditions where letter sequence is reversed from right to left. The effect was stronger for longer words (see Fig. 6). These findings support our hypothesis that readers first target the right part of the word strings and then moved the eyes backward, following the writing direction. Topical subheadings are allowed. 3 Discussion Natural text reading requires efficient coordination of cognitive and oculo- motor control processes. Cognitive principles are essential for the selection of an upcoming target word, however, it is the oculomotor system that provides the machinery to move the gaze to the region of the identified target word. In the current study, we set out to investigate whether ongoing cognition is able to overwrite default oculomotor control when the reader is confronted with manipulated (mirrored-reversed, inverted, and scambled) text layouts. On a global level, changing the display and positions of letters in a word modulates eye-movement statistics. Since mirrored-reversed and other ma- nipulated words are more difficult to process, a general tendency of longer average fixation durations and shorter mean saccade amplitudes is observed. These findings are in line with previous studies [19, 42]. Interestingly, sim- 20 Figure 6: Within-word landing position distribution for the initial landing position (first saccade) in two-fixation cases. 21 4−letter Words5−letter Words6−letter Words7−letter WordsMirrored Words (mW)Mirrored Letters (mL)Inverted Words (iW)Scrambled Letters (sL)02460246024602460.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.5Initial Landing PositionRelative Frequency Session:ControlSession 1Session 2Session 3Saccade 1 of Two−Fixation Cases ilar with findings from Kowler and Anton [19] and Kolers and Perkins [15], orthographical manipulations from our current study can be grouped into two categories based on their writing directions. Difficulties increased dra- matically, when readers read texts written against the reading direction as in mirrored words (mW) and inverted words (iW) conditions, where the letter- sequence of words is written from right to left within the displayed string due to mirroring or inversion. Our interpretation is supported by the fact that the observed deviations remained stable even after trainings. In contrast, during the manipulations that kept the letter-sequence in the reading direction (i.e., from left to right) in mirrored-letter word (mL), the patterns observed in eye movement measures (e.g., word-length effect) remained similar to those of normal reading and, after training, approached the behavior observed in normal reading. Finally, in our fourth experimental condition, we analyzed reading text composed of scrambled-letter (sL) words. Since letters in a word were randomly scrambled, there was no systematic way that the oculomotor behavior could adapt to process the new text layout. One of our motivations to run the current study was to obtain a detailed picture on within-word landing positions as the most important signature for oculomotor control. Mean landing position distribution of single fixations in reading manipulated texts does not significantly differ from reading normal text, although there was a slight shift observed in mirrored words (mW) and inverted words (iW) conditions. However, readers tend to increase their pre- cision in landing on word center when texts presentation was manipulated. The finding supports the idea that word center serves as saccade target lo- cation [5]. Analysis of mean landing position of forward saccades based on different launch sites demonstrated a general left shifts in experimental conditions compared to those observed in control condition. Especially in conditions where words are written against reading direction such as in mW and iW conditions, there is no clear preference for the eyes to initially land in the second half of the word strings (where the beginning of a word is located in this manipulation). This is contrary to our hypothesis. Interestingly, the peaks of the landing position in all manipulation types were shifted toward word centers, not necessaryly toward the beginning of word strings, compared to the peaks observed in reading normal text. In some rare undershoot cases in our data (e.g. launch site: -4, word length: 7 in iW and mW conditions), the peaks in experimental condition shift rightward of the peak observed in reading normal text. Furthermore, the variance of the initial landing position distribution was much smaller, meaning that the eyes landed more precisely 22 on the word center, which served as the theoretical target position. Given the precision of the landing position distributions and the highest likelihood to land on word center, we speculate that the leftward shifts were actually moving toward the word centers, not word beginning. Regarding the general improved precision of landing position distributions, a possible explanation is that with longer average fixation durations under our manipulations, the oculomotor system had more time to prepare the next saccade, which could result in a reduced saccadic error [28]. An alternative explanation is that unusual presentation of text was more salient and popped up parafoveally to enable precise saccadic targeting [12]. Additionally, modulation of manipulation types on launch-site effect was observed. In conditions where texts were written against reading direction (from right to left) we observed a reduced launch-site effect, while an in- creased launch-site effect was observed in conditions where texts were writ- ten normally. However, we are not sure if the difference effects are caused by the change in reading direction or increased processing difficulties. Nev- ertheless, the finding could also be explained in the theoretical framework of a Bayesian model of sensorimotor integration that we applied to reading [5, 21]. According to Bayes' rule, the observed landing position distribu- tion is the posterior distribution in a sensorimotor transform based on prior knowledge of typical target positions. If we assume that the prior knowledge is uninformative under unfamiliar text layout, then the posterior distribu- tion would shift toward the likelihood distribution, which is assumed to be unbiased with respect to the target word center. Therefore, in the Bayesian model, we could have expected a more centered landing-position distribution if we assumed that the experimental manipulations induce less prior knowl- edge on the possible target positions. Our finding demonstrated that by simply changing the letter-sequence information of words, we could obtain estimations for the slope parameter λ other than the typical value of 0.5. As a consequence, reading models development should aim for integrating a process-oriented model in generating saccade lengths. The most striking finding in the current study is the effect of text-display manipulation on initial landing position of two-fixation cases. When a word shows reversed letter ordering (hence a change in their spatial information) as a consequence of the experimental manipulation, we observed a clear shift into the second half of the word string when two fixations were generated. This finding demonstrates that the oculomotor system is able to adapt to display changes. Given that our hypothesis requires more than two fixations, 23 analysing the initial landing positions in refixation cases is more reasonable to test the hypothesis. Furthermore, our study also demonstrates that the typical usage of mean fixation position as a dependent variable to characterize oculomotor control is not specific enough to characterize control processes underlying reading. Since reading is a complex process, data delivered are usually complicated and require various analysis methods and inferences. Further systematic analyses and advanced mathematical modeling are required to investigate the dynamical processes underlying principles of oculomotor control and their interaction with ongoing cognition during reading. 4 Methods Participants A group of 37 participants (27 females, 10 males), aged between 16 and 39 years, received a total of EUR 70 for taking part in four 45-minute lab sessions and six 30-minute training sessions. They were all naive with respect to the purpose of the experiment. Participants reported normal or corrected- to-normal vision and declared their informed consent, p. The experiment conformed to the Declaration of Helsinki. Informed consent was obtained for experimentation by all participants. According to the standards of Deutsche Forschungsgemeinschaft (German Research Foundation) and German Society for Psychological Research, ethics committee approval was not required for this study. Apparatus, Material & Procedure Participants were assigned to four different groups based on four types of text manipulation, namely mirrored-word text (mW), mirrored-letter text (mL), inverted-word text (iW), and scrambled-word text (sL). Each participant did a total of four lab sessions and six training sessions at home via web-based interface. For lab sessions, participants read an excerpt from the German version of the novel "The Adventure of the Empty House" [? ]. They were seated at a viewing distance of 70cm in front of a 19-inch Mitsubishi Diamond Pro 2070 Monitor (screen resolution 1,280×1,024 pixels, refresh rate 100 Hz) 24 with the head supported by a chin rest. The stimuli (Courier font, size 18, black) were presented on vertical center line of the computer display with gray background color. Eye movements from both eyes were recorded using an EyeLink 1000 System (SR Research, Osgoode/Ontario, Canada) with a sampling rate of 1000 Hz and spatial resolution better than 0.01◦. At the end of the session, participants had to answer three questions related to the text they have just read. A maximum of 600 lines of text were presented over all lab sessions. For training sessions, participants read an excerpt from the novel "Small World" [48] by visiting a website created for the experiment1. After logging in, participants could read the manipulated text, which was presented as a line of max. 85 characters at the center of the screen, without time limit. When finished, they could go to the next line by clicking the right arrow or return to the previous page by clicking the left arrow. One training session should last at least 30 minutes. After logging out, participants received three questions via E-mail, which should be answered as soon as possible. No limit of the number of lines presented in training sessions was enforced. The complete procedures of the experiment went as follows: During the first lab session, participants read normal text (a total of 150 lines of maxi- mum 85 characters). After a two-hour break, participants read manipulated text in the second session, which lasted up to 45 minutes or when 150 lines were read. At home, participants were required to read manipulated text for two 30-minute sessions on the website before the third lab session. At the third lab session, participants continue reading manipulated text from where they left off at the previous session. Participants should conduct four 30-minute training sessions before taking part in the last lab session. Data Preparation Data containing blinks were discarded from the analysis. Saccades and fix- ations were detected using a velocity-based algorithm developed by Engbert et al. [9]. As a result, a total of 380, 292 fixations were detected. From this data set we excluded fixations based on the following criteria (i) fixations on the first and last words of a sentence as well as the first and the last of participant's fixation sequence, (ii) fixations with duration less than 20 ms or longer than 1000 ms, fixations landing outside the text rectangle, and sac- 1Using ShinyApps by RStudio accessible via http://www.shinyapps.io 25 cades shorter than one character space (12 pixels) or longer than 25 character spaces were removed from the analysis. Trials containing fixation duration longer than 2000 ms and less than three fixation points after filtering were ex- cluded from analyis. The remaining 236, 937 fixations are the valid fixations (see Table A1). Summary statistics In a global analysis, we computed statistics of fixation durations and saccade lengths of all valid fixations. For local processing, fixation duration and probability were grouped based on word length. Word lengths were grouped into three classes: short (≤ 4 characters), medium (5-7 characters) and long (> 7 characters) for further statistical analysis. Fixation durations. Visual information processing in reading is marked by the time spent on fixating words. Statistical analysis on single fixation For complimentary analysis on possible differences in processing All cases in which a word received exactly one fixation generates single fixation duration (SFD). In the cases in which a word received more than one fixations, the duration of the first and the second of those fixations were calculated for evaluating the first of multiple fixations (FMD) and second of multiple fixa- tions (SMD). All of those fixation duration measures were conducted on the first-pass reading data set. The sum of all fixations on a word, regardless the saccadic types, generated the total viewing time (TVT). For further statisti- cal analysis, the fixation duration measures on each word were transformed into their logarithmic values. Fixation probability. As measures of fixation probability we computed skipping probability (SKP), single fixation probability (SFP), and refixation probability (RFP) on first-pass reading data. Additionally, we calculated regression probability (RGP). For further statistical analysis, each word was assigned a logical value of 1 if it is skipped (SKP), fixated only once (SFP), fixated more than once (RFP) or a target of regressive saccades (RGP). Otherwise, a logical value of 0 was assigned to the word. To capture the changes of the manipulation effects on word processing difficulties across sessions, linear-mixed model analysis (3x4 factors) was conducted on the dependent variables single fixation duration (SFD) and refixation probability (RFP) as a function fixed word lenght effect and ran- dom effects generated by participants and sentences. The model was con- trasted against the word length effect during normal reading (sess = 0). We 26 used the lme4 package for R-Language of Statistical Computing [1] for esti- mating fixed (word length classes) and random (participant and sentence) coefficients. Specifically, each fixation duration and probability measure is modeled as a function of fixed word length effect (WLC across sessions (sess), with a fully parameterized variance-covariance matrix for participant (1 + WLC + sesspID) assuming participants may generate different slopes for each word length and sentence/line number (1sID) with the assumption that the intercept for each sentence do not vary accross sessions and word lengths. These model used the default treatment contrast. For example, the log-transformed measure of single fixation duration (SFD) is modeled as lmer(log(SFD) = WLC ∗ sess + (1 + WLC + sesspID) + (1sID)), (6) while refixation probability (RFP) are modeled as glmer(RFP = WLC ∗ sess + (1 + WLC + sesspID) + (1sID)), (7) because the glmer() function allows the statistical analysis of binary out- come. The models were estimated based on restricted maximum likelihood (REML). Furthermore, a 3x5-factor model using helmert constrast was con- ducted separately for each dependent measure with word lengths and manip- ulation types (type) as fixed main effects to investigate the different effect sizes among different manipulation types. The predictor type for reading normal text was used as baseline (type = 0). Initial Landing Position Within-word landing-position distributions in reading are broad and over- lap with neighboring words. As a consequence, observations of word-based landing positions in a reading experiment are truncated at word boundaries [28, 7]. In order to derive estimates of the means and standard deviations of the landing-position distribution, truncated normal curves for the distribu- tions of within-word fixation positions were fitted using Bayesian parameter inference [23] provided by the package rjags [35] in the R environment. Here we considered only cases where readers made forward saccades. Fix- ation data for each experimental condition and each experimental session were grouped based on word length and launch-site distance leading to 16 different word-length and launch-site specific data subsamples per condition and reading session. For each data subset Si, we estimated a two-dimensional 27 posterior distribution over the parameters mean µ and standard deviation σ of the underlying Gaussian landing-position distribution according to p(Siµ, 1/σ2)p(µ, 1/σ2) (cid:82)(cid:82) p(Siµ, 1/σ2)p(µ, 1/σ2) dµd1/σ2 . p(µ, 1/σ2Si) = (8) We assumed that observations of landing positions are generated by a normal- density likelihood function p(Siµ, 1/σ2) with mean µ and precision τ = 1/σ2. Furthermore, with p(µ, 1/σ2) we specified a normally distributed prior on the mean µ with mean M and precision T and a prior on τ distributed as a gamma density distribution with shape parameter A and rate B (see [23]). The parameters M, T and A, B for the prior over µ and τ, resp., were derived from the data in the following steps: For the control condition, landing- position distributions for each word length and launch-site distance were independently fitted by a truncated Gaussian function and the parameters of these fits were used to estimate the parameters of the empirical prior distribution. For reading manipulated text we also used the parameters of the prior of the control condition for estimating landing distributions of the first reading session and updated the prior for the later sessions systematically based on the results of the previous reading session. 5 References [1] Douglas Bates, Martin Mächler, Ben Bolker, and Steve Walker. Fitting linear mixed-effects models using lme4. Journal of Statistical Software, 67(1):1 -- 48, 2015. [2] Ben Bolker. bbmle: Tools for General Maximum Likelihood Estimation, 2017. R package version 1.0.20. [3] Christian Coëffé and J. Kevin O'Regan. Reducing the influence of non- target stimuli on saccade accuracy: Predictability and latency effects. Vision Research, 27(2):227 -- 240, 1987. [4] H. Deuble, W. Wolf, and G. Hauske. The evaluation of the oculomotor error signal. In Alastair G. Gale and Frank Johnson, editors, Theoretical and Applied Aspects of Eye Movement Research, volume 22 of Advances in Psychology, pages 55 -- 62. North-Holland, 1984. 28 [5] Ralf Engbert and André Krügel. Readers use bayesian estimation for eye movement control. Psychological Science, 21(3):366 -- 371, 2010. [6] Ralf Engbert, André Longtin, and Reinhold Kliegl. A dynamical model of saccade generation in reading based on spatially distributed lexical processing. Vision Research, 42(5):621 -- 636, 2002. [7] Ralf Engbert and Antje Nuthmann. Self-consistent estimation of mislo- cated fixations during reading. PLOS ONE, 3(2):1 -- 6, 02 2008. [8] Ralf Engbert, Antje Nuthmann, Eike M. Richter, and Reinhold Kliegl. Swift: A dynamical model of saccade generation during reading. Psy- chological Review, 112(4):777 -- 813, 2005. [9] Ralf Engbert, Petra Sinn, Konstantin Margenthaler, and Hans Truken- bord. Microsaccade toolbox. [10] J. M. Findlay. Global visual proceeding for saccadic eye movements. Vision Research, 22(8):1033 -- 1045, 1982. [11] John M Findlay and Iain D Gilchrist. Active vision: The psychology of looking and seeing. Oxford University Press, 2003. [12] Jukka Hyönä. Do irregular letter combinations attract readers' atten- tion? evidence from fixation locations in words. Journal of Experimental Psychology: Human Perception and Performance, 21(1):68 -- 81, Febru- ary 1995. [13] Rebecca L. Johnson and Emma L. Starr. The preferred viewing lo- cation in top-to-bottom sentence reading. The Quarterly Journal of Experimental Psychology, 71(1):220 -- 228, 2017. [14] Reinhold Kliegl, Ellen Grabner, Rolfs Martin, and Ralf Engbert. Length, frequency, and predictability effects of words on eye movements in read- ing. European Journal of Cognitive Psychology, 16(1 -- 2):262 -- 284, 2004. [15] P. A. Kolers and D. N. Perkins. Spatial and ordinal components of form perception and literacy. Cognitive Psychology, 7(2):228 -- 267, 1975. [16] Konrad P. Körding. Decision theory: What "should" the nervous system do? Science, 318(5850):606 -- 610, 2007. 29 [17] Konrad P. Körding and Daniel M. Wolpert. Bayesian integration in sensorimotor learning. Nature, 427(6971):244 -- 247, 2004. [18] Konrad P. Körding and Daniel M. Wolpert. Bayesian decision theory in sensorimotor control. Trends in Cognitive Sciences, 10(7):319 -- 326, 2006. [19] Eileen Kowler and Stanley Anton. Reading twisted text: Implications for the role of saccades. Vision Research, 27(1):45 -- 60, 1987. [20] André Krügel and Ralf Engbert. On the launch-site effect for skipped words during reading. Vision Research, 50(16):1532 -- 1539, 2010. [21] André Krügel and Ralf Engbert. A model of saccadic landing positions in reading under the influence of sensory noise. Visual Cognition, 22(3- 4):37 -- 41, 2014. [22] André Krügel, Françoise Vitu, and Ralf Engbert. Fixation positions af- ter skipping saccades: A single space makes a large difference. Attention, Perception, & Psychophysics, 74(8):1556 -- 1561, 2012. [23] John K. Kruschke. Doing Bayesian data analysis: A tutorial with R, JAGS, and Stan. Academic Press, 2014. [24] Frederic Lavigne, Francoise Vitu, and Gery d'Ydewalle. The infuence of semantic context on initial eye landing sites in words. Acta Psychologia, 104(191-214), 2000. [25] Gordon E. Legge, Thomas A. Hooven, Timothy S. Klitz, J. Stephen Mansfield, and Bosco S. Tjan. Mr. chips 2002: new insights from an ideal-observer model of reading. Vision Research, 42(18):2219 -- 2234, 2002. [26] Gordon E. Legge, Timothy S. Klitz, and Bosco S. Tjan. Mr. chips: An ideal-observer model of reading. Psychological Review, 104(3):524 -- 553, 1997. [27] Steven G Luke and John M Henderson. Oculomotor and cognitive con- trol of eye movements in reading: evidence from mindless reading. At- tention, Perception, & Psychophysics, 75(6):1230 -- 1242, 2013. 30 [28] G. W. McConkie, P. W. Kerr, M.D. Reddix, and D Zola. Eye movement control during reading: I. the localisation of initial eye fixations in words. Visual Research, 28(10):1107 -- 1118, 1988. [29] Antje Nuthmann. The "where" and "when" of eye fixations in reading. Unpublished Doctoral Dissertation, University of Potsdam, Germany (online available from ttp://nbn-resolving.de/urn:nbn:de:kobv:517- opus-7931), 2006. [30] Antje Nuthmann, Ralf Engbert, and Reinhold Kliegl. Mislocated fix- ations during reading and the inverted optimal viewing position effect. Vision Research, 45(17):2201 -- 2217, 2005. [31] Antje Nuthmann, Ralf Engbert, and Reinhold Kliegl. The iovp ef- fect in mindless reading: Experiment and modeling. Visual Research, 47(7):990 -- 1002, 2007. [32] J. Kevin O'Regan. Eye movements and reading. In E. Kowler (Ed.), Eye movements and their role in visual and cognitive processes (pp. 395- 453). Elsevier, Amsterdam, 1990. [33] J. Kevin O'Regan and A. Levy-Schoen. Eye movement strategy and tactics in word recognition and reading. In M.Coltheart (Ed.), Attention and performance XII: The psychology of reading(pp. 363-383). Erlbaum, Hillsdale, NJ, 1987. [34] J. Kevin O'Regan, A. Levy-Schoen, J. Pynte, and B. Brugaillere. Con- venient fixation location within isolated words of different length and structure. Journal of Experimental Psychology: Human Perception and Performance, 10(2):250257, 1984. [35] Martyn Plummer. rjags: Bayesian Graphical Models using MCMC, 2016. R package version 4-6. [36] E. C. Poulton. Human manual control. In Supplement 2: Handbook of Physiology, The Nervous System, Motor Control, pages 1337 -- 1389. Wiley-Blackwell, 1981. [37] Ralph Radach. Blickbewegungen beim Lesen: Psychologische Aspekte der Determination von Fixationspositionen. Waxmann Verlag, 1996. 31 [38] Ralph Radach and G. W. McConkie. Determinants of fixation positions in words during reading. Eye Guidance in Reading and Scene Perception, pages 77 -- 100, 1998. [39] Keith Rayner. Eye guidance in reading: Fixation locations within words. Perception, 8(1):21 -- 30, 1979. [40] Keith Rayner. Eye movements and attention during reading, scene per- ception, and visual search. The Quarterly Journal of Experimental Psy- chology, 62(8):1457 -- 1506, 2009. [41] Keith Rayner, K. S. Binder, J. Ashby, and Alexander Pollatsek. Eye movement control in reading: word predictability has little influence on initial landing positions in words. Vision Research, 41(7):943 -- 54, 2001. [42] Keith Rayner, Sarah J. White, Rebecca L. Johnson, and Simon P. Liv- ersedge. Raeding wrods with jubmled lettres: There is a cost. Psycho- logical Science, 17(3):192 -- 193, 2006. [43] E. D. Reichle, Alexander Pollatsek, and Keith Rayner. E -- z reader: A cognitive-control, serial-attention model of eye-movement behavior during reading. Cognitive Systems Research, 7(1):4 -- 22, 2006. [44] Erik D. Reichle, Alexander Pollatsek, Donald L. Fischer, and Keith Rayner. Toward a model of eye movement control in reading. Psycho- logical Review, 105(1):125 -- 157, 1998. [45] Erik D. Reichle, Keith Rayner, and Alexander Pollatsek. Eye movement control in reading: accounting for initial fixation locations and refixa- tions within the e-z reader model. Vision Research, 39(26):4403 -- 4411, 1999. [46] Erik D. Reichle, Tessa Warren, and Kerry McConnell. Using e-z reader to model the effects of higher level language processing on eye movements during reading. Psychonomic Bulletin & Review, 16(1):1 -- 21, 2009. [47] Daniel Schad and Ralf Engbert. The zoom lens of attention: Simulat- ing shuffled versus normal text reading using the swift model. Visual Cognition, 20(4-5):391 -- 421, 04 2012. [48] Martin Suter. Small World. Diogenes, Zürich, 1997. 32 [49] Françoise Vitu. The existence of a center of gravity effect during reading. Vision Research, 31(7-8):1289 -- 1313, 1991. [50] Françoise Vitu. About the global effect and the critical role of retinal eccentricity: Implications for eye movements in reading. Journal of Eye Movement Research, 2(3):1 -- 18, 2008. [51] Françoise Vitu, D. Lancelin, A. Jean, and F. Farioli. Influence of foveal distractors on saccadic eye movements: A dead zone for the global effect. Vision Research, 46(28):4684 -- 4708, 2006. [52] Francoise Vitu, George W. McConkie, Paul Kerr, and J. Kevin O'Regan. Fixation location effects on fixation durations during reading: an in- verted optimal viewing position effect. Vision Research, 41(25-26):3513 -- 3533, 2001. [53] Francoise Vitu, J. Kevin O'Regan, and M. Mittau. Optimal landing position in reading isolated words and continuous text. Perception & Psychophysics, 47(6):583 -- 600, 1990. [54] Sarah J. White, Rebecca L. Johnson, Simon P. Liversedge, and Keith Rayner. Eye movements when reading transposed text: The importance of word-beginning letters. Journal of Experimental Psychology: Human Perception and Performance, 34(5):1261 -- 1276, 2008. [55] Sarah J. White and Simon P. Liversedge. Orthographic familiarity in- fluences initial eye fixation positions in reading. European Journal of Cognitive Psychology, 16(1-2):52 -- 78, 2004. [56] Sarah J. White and Simon P. Liversedge. Foveal processing does not modulate non-foveal orthographic influences on fixation positions. Vi- sion Research, 46(3):426 -- 437, 2006. [57] Sarah J. White and Simon P. Liversedge. Linguistic and nonlinguistik influences on the eyes' landing positions during reading. The Quarterly Journal of Experimental Psychology, 59(4):760 -- 782, 2006. [58] Daniel M. Wolpert and Zoubin Ghahramani. Bayes rule in perception, action and cognition. In The Oxford Companion to the Mind, pages 1 -- 4. Oxford University Press, 2005. 33 6 Acknowledgements This work was supported by Deutsche Forschungsgemeinschaft (grant EN 471/15-1 to R. E.). 7 Author contributions statement R.E. and A.K. developed the study concept and the study design. Testing and data collection were performed by J.C.. J.C. and A.K. performed the data analysis and all authors contributed to the interpretation. J.C. and A.K. drafted the manuscript, and R.E. provided critical revisions. All authors approved the final version of the manuscript for submission. 8 Additional information Competing interests The authors declare no competing interests. A Appendix 34 ) L s ( s r e t t e L d e l b m a r c S ) W i ( s d r o W d e t r e v n I . s c i t s i t a t s y r a m m u s l a b o l G : 1 A e l b a T ) L m ( s r e t t e L d e r o r r i M ) W m ( s d r o W d e r o r r i M l o r t n o C n o i t i d n o C 3 2 6 0 4 2 2 8 4 6 6 2 1 9 4 5 5 2 3 8 2 7 5 2 2 4 0 1 4 2 1 1 3 5 2 2 3 6 4 2 2 2 2 5 0 2 4 2 1 7 7 6 6 2 3 7 5 9 6 2 2 0 0 5 7 2 1 8 9 5 1 2 0 7 8 4 2 8 N 7 9 5 0 1 4 0 . 4 4 5 6 4 3 1 6 9 . 5 5 6 3 8 8 2 6 . 5 6 3 4 8 3 9 4 . 3 4 7 0 8 1 5 4 . 0 2 0 7 1 3 8 8 . 5 3 6 6 9 3 9 . 0 1 4 0 3 0 5 3 1 1 9 5 . 2 4 8 9 2 5 1 1 4 . 7 5 4 3 0 0 1 9 5 . 5 6 9 1 2 4 5 0 . 2 4 9 9 9 1 2 9 . 9 1 5 0 8 3 2 9 . 7 3 5 8 9 2 8 . 9 0 0 3 1 6 3 1 1 7 4 . 4 4 8 8 1 4 1 3 5 . 5 5 9 6 2 0 1 8 3 . 2 7 7 8 8 4 9 5 . 7 4 3 7 9 1 1 2 . 9 1 0 0 4 3 1 1 . 3 3 2 9 9 6 6 . 9 6 0 3 3 3 9 0 1 9 4 . 2 4 5 9 7 4 1 1 5 . 7 5 2 1 9 1 1 1 5 . 0 8 3 6 3 5 2 0 . 5 4 4 7 7 5 . 6 2 7 7 5 6 4 . 8 4 9 4 6 5 4 . 5 9 2 3 6 6 7 0 1 6 6 . 4 4 8 3 3 3 1 4 3 . 5 5 8 2 4 0 1 8 1 . 8 7 3 5 6 4 2 6 . 4 4 4 7 6 6 4 . 6 6 9 0 5 7 8 . 8 4 4 8 5 6 . 5 4 4 3 0 4 3 0 1 9 8 . 5 4 1 9 1 2 1 1 1 . 4 5 9 7 4 9 5 7 . 7 7 0 5 0 4 3 7 . 2 4 1 3 4 5 5 . 4 9 8 9 4 3 6 . 2 5 9 4 5 9 7 . 5 3 8 3 1 7 2 1 1 7 6 . 0 5 5 7 9 0 1 3 3 . 9 4 9 0 8 8 6 2 . 0 8 9 0 9 3 8 3 . 4 4 4 6 9 1 3 . 2 2 0 3 9 2 6 2 . 3 3 8 3 6 4 2 . 7 6 6 2 9 3 6 1 1 9 0 . 8 4 6 6 5 2 1 1 9 . 1 5 1 6 0 0 1 7 0 . 0 8 9 7 3 4 2 5 . 3 4 3 9 1 2 8 . 1 2 6 8 4 3 5 6 . 4 3 5 9 6 1 9 . 6 7 7 2 3 4 1 2 1 2 5 . 5 4 4 3 5 4 1 8 4 . 4 5 0 0 2 1 1 6 0 . 7 7 6 3 6 5 2 3 . 0 5 6 7 4 1 8 1 . 3 1 2 8 0 4 5 4 . 6 3 4 3 7 5 5 . 6 9 9 2 1 9 6 1 1 7 3 . 3 4 6 6 2 5 1 3 6 . 6 5 0 2 6 1 1 2 1 . 6 7 5 2 7 5 7 2 . 9 4 1 3 9 1 0 . 8 7 5 9 4 6 6 . 2 4 4 5 8 5 3 . 7 7 3 3 1 5 2 2 1 5 5 . 4 4 9 4 2 5 1 5 4 . 5 5 6 6 1 1 1 2 2 . 3 7 6 8 3 5 4 2 . 8 4 4 3 8 7 4 . 7 3 3 9 4 8 1 . 4 4 8 0 8 4 2 . 7 1 5 3 1 5 3 0 1 3 9 . 7 4 7 4 2 1 1 7 0 . 2 5 5 2 9 6 7 5 . 1 6 0 6 9 2 4 7 . 2 4 6 1 5 5 4 . 7 6 3 4 3 2 6 . 9 4 3 9 6 1 0 . 0 1 4 5 3 7 7 3 0 4 5 9 . 8 4 0 1 1 2 4 5 0 . 1 5 3 4 3 2 3 1 8 . 6 7 6 5 9 4 1 4 2 . 6 4 3 6 9 9 8 . 0 3 7 5 3 7 5 7 . 2 2 2 9 6 3 2 4 . 1 1 5 4 2 8 3 . 6 5 1 . 6 8 1 . 6 1 8 . 5 5 5 . 5 5 2 . 5 6 8 . 6 6 5 . 6 3 5 . 5 1 8 . 5 8 5 . 5 9 1 . 5 2 8 . 7 N % N % N % N % N % N % N % n a e m ] s m [ n a e m ] r a h c [ 5 2 . 4 5 9 . 3 6 6 . 4 9 7 . 2 7 5 . 2 3 6 . 2 7 6 . 2 8 8 . 2 1 0 . 3 3 0 . 3 7 9 . 2 2 8 . 2 3 2 . 4 n a e m ] r a h c [ d e t c e t e D s n o i t a x fi n o i s s e S d e d r a c s i D n o i t a x fi s s a p - t s r i F n o i t a x fi d i l a V d r a w r o F s e d a c c a s i g n p p i k S s e d a c c a s n o i t a x fi s n o i t a x fi e R s n o i s s e r g e R n o i t a r u d n o i t a x fi d r a w r o f e d a c c a s e d u t i l p m a d r a w k c a b e d a c c a s e d u t i l p m a 35 Table A2: Results from Linear Mixed-Effects: Single Fixation Duration Condition Mirrored Letters (mL) Mirrored Words (mW) Inverted Words (iW) Scrambled Letters (sL) Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 b 5.47 0.02 0.18 0.10 0.09 0.06 0.07 0.06 5.44 0.02 0.31 0.31 0.29 -0.14 -0.07 -0.01 5.48 0.01 0.36 0.27 0.22 -0.21 -0.10 -0.06 5.46 0.01 0.11 0.11 0.10 0.10 0.12 0.09 SE 0.04 0.02 0.02 0.02 0.02 0.01 0.01 0.01 0.04 0.02 0.07 0.06 0.06 0.01 0.01 0.01 0.04 0.02 0.05 0.03 0.03 0.01 0.01 0.01 0.03 0.02 0.04 0.05 0.04 0.01 0.01 0.01 t value 124.53 1.15 9.69 5.32 4.22 7.79 9.38 8.30 142.66 0.87 4.54 5.11 5.10 -12.13 -7.08 -0.99 128.67 0.45 7.48 7.84 6.17 -18.10 -9.56 -5.71 186.58 0.24 2.91 2.21 2.34 11.95 14.61 10.80 Note: Fixation duration value is log-transformed. WLC = word length. Non-significant values are marked in bold. 36 Table A3: Results from Linear Mixed-Effects model: Duration of First of Multiple Fixations Conditions Mirrored Letters (mL) b 5.43 -0.01 0.16 0.04 0.03 0.09 0.09 0.08 5.37 0.04 0.34 0.32 0.24 -0.16 -0.20 -0.10 5.46 0.03 0.33 0.25 0.20 -0.16 -0.14 -0.11 5.44 0.01 0.06 0.06 0.06 0.15 0.22 0.15 SE 0.05 0.02 0.03 0.03 0.03 0.01 0.02 0.02 0.05 0.03 0.04 0.03 0.05 0.02 0.01 0.01 0.04 0.02 0.05 0.05 0.05 0.02 0.02 0.02 0.03 0.02 0.04 0.03 0.03 0.02 0.02 0.02 t value 114.86 -0.31 4.56 1.23 0.93 6.04 5.86 4.68 101.59 1.56 9.23 10.98 4.36 -10.31 -14.05 -6.60 122.19 1.22 6.32 4.83 4.14 -10.58 -9.11 -7.32 162.02 0.27 1.54 1.84 1.77 8.80 12.83 8.52 Mirrored Words (mW) Inverted Words (iW) Scrambled Letters (sL) Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Note: Fixation duration value is log-transformed. WLC = word length. Non-significant values are marked in bold. 37 Table A4: Results from Linear Mixed-Effects model: Refixation Probabilities Condition Mirrored Letters (mL) b Mirrored Words (mW) Inverted Words (iW) Scrambled Letters (sL) Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 Intercept WLC Session 1 Session 2 Session 3 WLC:Session 1 WLC:Session 2 WLC:Session 3 -1.63 1.15 1.29 0.76 0.48 0.37 0.41 0.38 -1.35 1.15 1.39 1.35 1.19 0.55 0.64 0.69 -1.51 1.21 1.87 1.69 1.65 0.64 0.52 0.62 -1.51 1.46 0.78 0.57 0.67 0.32 0.42 0.31 SE 0.12 0.09 0.18 0.14 0.17 0.05 0.06 0.06 0.18 0.11 0.23 0.14 0.21 0.05 0.05 0.05 0.17 0.07 0.15 0.17 0.17 0.05 0.05 0.05 0.15 0.10 0.23 0.20 0.18 0.05 0.06 0.06 z value Pr(>z) -13.75 12.52 7.32 5.53 2.75 6.86 7.33 6.48 -7.62 10.50 6.15 9.78 5.71 10.00 12.98 13.85 -9.03 18.14 12.49 10.06 9.98 11.76 9.96 12.17 -9.73 13.95 3.48 2.90 3.64 5.92 7.61 5.52 0.000 0.000 0.000 0.000 0.006 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.004 0.000 0.000 0.000 0.000 Note: WLC = word length. Non-significant values are marked in bold. 38 b 5.65 Intercept -0.03 WLC 0.09 mL 0.01 sL 0.05 mW 0.04 iW 0.03 WLC:mL WLC:sL 0.02 WLC:mW -0.05 WLC:iW -0.04 5.60 Intercept 0.00 WLC mL 0.08 -0.01 sL 0.06 mW 0.04 iW WLC:mL 0.04 WLC:sL 0.05 WLC:mW -0.06 WLC:iW -0.04 Intercept -0.40 1.59 0.62 mL 0.04 sL 0.21 mW 0.21 iW 0.17 WLC:mL WLC:sL 0.08 WLC:mW 0.07 WLC:iW 0.05 SE 0.03 0.01 0.02 0.02 0.01 0.01 0.00 0.00 0.00 0.00 0.03 0.01 0.02 0.01 0.01 0.01 0.01 0.01 0.00 0.00 0.08 0.04 0.08 0.06 0.04 0.04 0.02 0.02 0.01 0.01 z value 213.36 -3.57 3.95 0.46 3.90 4.07 6.97 8.38 -19.63 -19.74 212.41 -0.29 4.08 -0.42 5.47 4.45 5.82 9.21 -17.12 -14.35 -5.03 36.17 7.33 0.66 4.80 5.91 7.17 4.47 5.05 4.54 First of Multiple Fixations duration (FMF) Refixation Probability WLC (RFX) Table A5: Results from Linear Mixed-Effects model: Effect sizes accross manipulation types Measures Single Fixation Duration (SFD) Pr(>z) 0.000 0.000 0.000 0.506 0.000 0.000 0.000 0.000 0.000 0.000 Note: LMM model with helmert contrast. SFD and FMD values are log-transformed (base 10). WLC = word length. Non-significant values are marked in bold. 39 Table A6: Quadratic fit to initial landing position curve (two-fixation cases) for reading normal text: Estimates of parameters A, B and C Session Word Length Parameters Estimate Std. Error 0 4 4 4 5 5 5 6 6 6 7 7 7 0.020 0.008 0.101 0.015 0.004 0.121 0.012 0.003 0.168 0.016 0.003 0.251 z value Pr(z) 0.013 2.466 0.000 7.814 24.916 0.000 0.134 1.497 0.000 9.139 0.000 27.468 1.516 0.130 0.000 8.859 0.000 25.043 -0.518 0.604 0.000 7.115 19.834 0.000 A B C A B C A B C A B C 0.051 0.065 2.528 0.023 0.040 3.336 0.019 0.023 4.21 -0.008 0.018 4.98 40 Table A7: Quadratic fit to initial landing position curve (two-fixation cases) for reading mirrored-word (mW) text: Estimates of parameters A, B and C Session Word Length Parameters Estimate Std. Error 1 2 3 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C 0.080 0.048 2.670 0.144 0.007 2.787 0.181 -0.009 3.523 0.173 -0.009 3.267 0.074 0.019 4.152 0.099 0.012 4.202 0.056 0.000 17.788 0.166 -0.007 3.057 0.043 0.063 2.718 0.0845 0.011 4.708 -0.008 0.001 17.752 0.176 -0.009 3.101 41 0.022 0.009 0.173 0.027 0.007 0.757 0.028 0.005 0.602 0.019 0.003 0.314 0.039 0.010 1.145 0.027 0.008 1.221 0.086 0.000 0.000 0.025 0.004 0.526 0.011 0.005 0.069 0.042 0.009 2.078 0.076 0.000 0.001 0.028 0.004 0.465 z value 3.684 5.309 15.650 5.288 1.037 3.684 6.533 -1.680 5.852 9.009 -3.265 10.392 1.897 1.9497 3.627 3.633 1.518 3.443 0.650 1.049 38076.98 6.692 -2.078 5.806 3.861 13.502 39.444 1.998 1.119 2.265 -0.101 2.050 6.275 -2.315 6.666 23601.67 Pr(z) 0.000 0.000 0.000 0.000 0.300 0.000 0.000 0.093 0.000 0.000 0.001 0.000 0.058 0.051 0.000 0.000 0.129 0.001 0.516 0.294 0.000 0.000 0.038 0.000 0.000 0.000 0.000 0.046 0.263 0.023 0.919 0.040 0.000 0.000 0.021 0.000 Table A8: Quadratic fit to initial landing position curve (two-fixation cases) for reading mirrored-letter (mL) text: Estimates of parameters A, B and C Session Word Length Parameters Estimate Std. Error 1 2 3 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C -0,007 0,078 2,802 0,009 0,038 3,593 0,028 0,016 4,800 0,015 0,007 6,717 0,002 0,060 3,150 1,316 -0,002 -20,033 0,002 0,022 4,544 0,011 0,010 5,933 0,041 0,054 2,980 0,010 0,037 3,654 0,002 0,024 4,363 -0,002 0,017 4,990 42 0,022 0,009 0,120 0,017 0,005 0,172 0,006 0,001 0,161 0,016 0,002 0,787 0,004 0,002 0,035 0,331 0,001 0,018 0,019 0,004 0,324 0,010 0,002 0,405 0,023 0,010 0,211 0,021 0,006 0,227 0,020 0,004 0,290 0,020 0,003 0,329 z value Pr(z) 0.753 -0,314 0.000 8,332 23,424 0.000 0.621 0,494 0.000 7,626 0.000 20,842 4,569 0.000 0.000 12,389 0.000 29,770 0,943 0.346 0.000 4,283 0.000 8,540 0,446 0.656 0.000 36,706 0.000 89,4066 0.000 3,974 -3,487 0.000 0.000 -1087,91 0.924 0.000 0.000 0.282 0.000 0.000 0.066 0.000 0.000 0.631 0.000 0.000 0.936 0.000 0.000 0.902 0.000 0.000 0,095 5,463 14,023 1,076 6,450 14,661 1,838 5,436 14,152 0,487 6,021 16,110 0,080 5,568 15,042 -0,123 5,455 15,169 Table A9: Quadratic fit to initial landing position curve (two-fixation cases) for reading inverted-word (iW) text: Estimates of parameters A, B and C Session Word Length Parameters Estimate Std. Error 1 2 3 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C -0.460 0.002 22.504 0.213 -0.012 1.510 0.214 -0.018 2.940 0.197 -0.013 3.064 0.058 0.030 3.677 0.006 0.002 11.017 0.042 0.000 24.665 0.167 -0.008 3.522 0.021 0.053 3.185 0.080 0.014 4.295 0.057 0.000 25.887 0.165 -0.008 3.685 43 0.116 0.000 0.004 0.022 0.006 0.653 0.026 0.005 0.243 0.023 0.003 0.276 0.014 0.005 0.309 0.054 0.001 0.001 0.116 0.000 0.001 0.025 0.004 0.439 0.010 0.004 0.106 0.029 0.008 1.099 0.114 0.000 0.000 0.022 0.003 0.426 z value -3.976 5.742 5642.92 9.688 -1.884 2.312 8.132 -3.570 12.093 8.550 -3.954 11.102 4.153 5.758 11.908 0.115 3.198 11956.68 0.361 0.881 41339.44 6.785 -2.279 8.024 2.219 12.494 30.049 2.767 1.763 3.908 0.502 0.767 7.515 -2.388 8.654 51952.32 Pr(z) 0.000 0.000 0.000 0.000 0.060 0.021 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.909 0.001 0.000 0.718 0.378 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.006 0.078 0.000 0.616 0.443 0.000 0.000 0.017 0.000 Table A10: Quadratic fit to initial landing position curve (two-fixation cases) for reading scrambled-letter (sL) text: Estimates of parameters A, B and C Session Word Length Parameters Estimate Std. Error 1 2 3 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 4 4 4 5 5 5 6 6 6 7 7 7 A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C A B C -0.034 0.087 2.833 0.017 0.031 3.861 0.022 0.019 4.563 0.620 0.000 -20.736 0.002 0.086 2.538 -0.009 0.042 3.580 0.021 0.014 5.149 0.618 -0.001 -20.966 1.376 -0.002 -22.215 0.025 0.029 3.918 0.020 0.015 5.046 -0.027 0.003 10.061 44 0.034 0.014 0.169 0.008 0.0024 0.118 0.008 0.002 0.155 0.043 7.177 0.001 0.026 0.010 0.095 0.024 0.007 0.210 0.015 0.003 0.492 0.055 0.000 0.001 0.758 0.001 0.041 0.009 0.003 0.145 0.020 0.004 0.601 0.011 0.000 0.000 z value -1.020 6.051 16.776 2.123 13.118 32.791 2.859 11.561 29.519 14.332 -11.598 -19706.55 0.081 8.373 26.578 -0.363 6.196 17.015 1.348 4.705 10.456 11.303 -9.141 -17086.51 1.815 -1.553 -547.09 2.778 10.999 27.020 0.984 3.698 8.399 -2.442 16.467 49984.09 Pr(z) 0.301 0.000 0.000 0.034 0.000 0.000 0.004 0.000 0.000 0.000 0.000 0.000 0.935 0.000 0.000 0.716 0.000 0.000 0.178 0.000 0.000 0.000 0.000 0.000 0.070 0.121 0.000 0.005 0.000 0.000 0.325 0.000 0.000 0.015 0.000 0.000 Figure A1: Within-word landing position distribution group by launch-site distance and word length for reading mirrored-word (mW) texts. Red line and dots represent data from normal reading session. Data from the first experimental session are presented in dark blue color. The lighter blue hues represent the last two experimental sessions. 45 3−letterwords4−letterwords5−letterwords6−letterwords7−letterwordslaunch site−5launch site−4launch site−3launch site−2launch site−102468024680246802468024680.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.5Initial Landing Position [char]Relative FrequencySession:ControlSession 1Session 2Session 3Landing Position Distribution: Mirrored Words (mW) Figure A2: Within-word landing position distribution group by launch-site distance and word length for reading mirrored-letter (mL) texts. Red line and dots represent data from normal reading session. Data from the first experimental session are presented in dark blue color. The lighter blue hues represent the last two experimental sessions. 46 3−letterwords4−letterwords5−letterwords6−letterwords7−letterwordslaunch site−5launch site−4launch site−3launch site−2launch site−102468024680246802468024680.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.5Initial Landing Position [char]Relative FrequencySession:ControlSession 1Session 2Session 3Landing Position Distribution: Mirrored Letters (mL) Figure A3: Within-word landing position distribution group by launch-site distance and word length for reading inverted-word (iW) texts. Red line and dots represent data from normal reading session. Data from the first experimental session are presented in dark blue color. The lighter blue hues represent the last two experimental sessions. 47 3−letterwords4−letterwords5−letterwords6−letterwords7−letterwordslaunch site−5launch site−4launch site−3launch site−2launch site−102468024680246802468024680.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.5Initial Landing Position [char]Relative FrequencySession:ControlSession 1Session 2Session 3Landing Position Distribution: Inverted Words (iW) Figure A4: Within-word landing position distribution group by launch-site distance and word length for reading scrambled-letter (sL) texts. Red line and dots represent data from normal reading session. Data from the first experimental session are presented in dark blue color. The lighter blue hues represent the last two experimental sessions. 48 3−letterwords4−letterwords5−letterwords6−letterwords7−letterwordslaunch site−5launch site−4launch site−3launch site−2launch site−102468024680246802468024680.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.50.00.10.20.30.40.5Initial Landing Position [char]Relative FrequencySession:ControlSession 1Session 2Session 3Landing Position Distribution: Scrambled Letters (sL)
1208.6041
3
1208
2012-10-03T19:21:32
Response Selection Using Neural Phase Oscillators
[ "q-bio.NC", "physics.bio-ph" ]
In a recent paper, Suppes et al. (2012) [arXiv:arXiv:1010.3063] used neural oscillators to create a model, based on reasonable neurophysiological assumptions, of the behavioral stimulus-response (SR) theory. In this paper, we describe the model in a less mathematical and more physical and intuitive way.
q-bio.NC
q-bio
Response Selection Using Neural Phase Oscillators J. Acacio de Barros∗, and G. Oas† ∗Liberal Studies, San Francisco State University, San Francisco, CA, 94132 †CSLI, Stanford University, Stanford, CA 94305 Abstract In a recent paper, Suppes, de Barros, & Oas (2012) used neural oscillators to cre- ate a model, based on reasonable neurophysiological assumptions, of the behavioral stimulus-response (SR) theory. In this paper, we describe the main characteristics of the model, emphasizing its physical and intuitive aspects. 1 Introduction It is an honor for Acacio de Barros and Gary Oas to participate in a festschrift for Pat Suppes. It is especially rewarding to do so with a paper where we discuss our most recent work with Pat, a model of brain processes using neural oscillators. We are, as Pat would say, "true blue physicists." So, for us, collaborating with Pat on this truly interdisciplinary paper is an example not only of his intellectual influence, but also of his friendship and mentorship. We are happy to dedicate this paper to Pat. Happy Birthday Pat! The work we present here started more than ten years ago, when JAB and Pat begun thinking about how to model in a physically plausible way collections of neurons in terms of oscillators. In one of his known intuitions, Pat kept insisting that the brain "gotta use oscillators." Of course, as is often the case, his "intuition" was based on hard work and detailed empirical data that he collected working on the EEG of words and sentences. Nevertheless, as we kept trying to make our model work (and we had many failures, and some successes; see Vassilieva et al. (2011) for an example), Pat kept insisting: we should understand the brain computations with oscillators. I am pleased to say that, despite my initial skepticism, we now have a model that we feel is not only grounded on neurophysiologically sound evidence, but that also reproduces quite well some empirical behavioral data (Suppes, de Barros, & Oas, 2012). In this paper we attempt to describe the main features this model by fo- cusing on the physical processes underlying the neural computations. We chose to do so for the following reasons. First, because of its interdisciplinarity, our 1 1 Introduction 2 model requires concepts from many different areas (neurophysiology, physics, psychology, etc). Such concepts are not complex, but are often unfamiliar to most researchers. Second, we are confident that our model is relevant to cogni- tive psychologists, as it may explain some mathematical models showing good empirical fit (de Barros, 2012a,b). So, we believe that this paper can provide a clearer and intuitive view of the main physical features of our model for those thinking about applying it, supplementing the discussions found in Suppes, de Barros, & Oas (2012). Let us start our discussion with the broad problem of understanding how the brain processes information. This is perhaps the most challenging cur- rent scientific endeavors, mainly due to the fact that our brain is tremendously complicated, as it is constituted of many different components that are, by themselves, complex, but that also seem to sometimes interact holistically with each other. Among the approaches to try and understand the brain, the most prominent ones are the top-down and bottom-up. In the top-down approach, we start with the higher-level functions and go to their underlying mechanisms. An example of such approach would be the field of cognitive neuroscience, where often one starts with experiments in cognitive psychology and tries to under- stand them from principles in neuroscience (Adolphs, 2003). In the bottom-up approach, one tries to start with neurophysiology, and by studying how each elementary component works, one tries to see how higher functions arise from such components or their interaction (Kandel et al., 2000). Each of those approaches have their shortcomings. For example, one of the main issues is what we may call a problem of scale. When trying to understand a complex system, the first question that arises is how detailed we need to be. In the case of the brain, some researchers say that we need to go all the way down to the chemical reactions in the synapses. Others argue that individual neurons hold the key to understanding brain computation. Yet another view is that collections of neurons are important. So, when trying to understand how the brain works, our first problem is where to begin. Regardless of what scale is chosen and where we start, ultimately we would need to understand the whole process if we were to claim to have understood the brain. The main problem with connecting a higher scale with a lower one is due to its complexity. For example, evidence exists that higher cognitive processes involve tens to hundreds of thousands of neurons, interacting with each other in very complex ways. Modeling such processes require the use of powerful com- puters. But, even when a model is shown to work from the underlying neuronal dynamics, the use of massive computer simulations helps little in understanding, in an intuitive or conceptual way, what is actually happening. The system is simply too complex. To deal with the issue of complexity, different approaches can be taken. One possible route is to find physically plausible arguments that impose constraints on the system's dynamics, therefore reducing it to fewer degrees of freedom. This is the approach taken by Suppes, de Barros, & Oas (2012). In their paper, a large number of independent neurons was modeled by a single dynamical parameter determined by the phase of a neural oscillator. They then showed that under 2 A Brief Review of SR theory 3 certain reasonable assumptions, the main characteristics of behavioral stimulus- response (SR) theory could be described by neural oscillators. The use of neural oscillators thus provided a significant reduction on the number of degrees of freedom, allowing for the physical interpretation of many different parameters in the model. In this paper we present the work of Suppes, de Barros, & Oas (2012), with emphasis on the physics and intuition behind the model. Our goal is to make this model more understandable, as many of the concepts used in our previous paper are not well-known to certain audiences. For example, while all physicists have an excellent knowledge of oscillations and interference and could easily follow the arguments leading from neurons to oscillators, only a few would feel comfortable with the mathematical learning theories used. Neuroscientists, on the other hand, would probably feel at home with neurons and learning theories, but not so much with oscillators and interference. Neither would most psychologists. Here we focus on the intuitions behind the physics, with the hopes that, in conjunction with the oscillator model, psychologists and neuroscientists could benefit more from the insights gained. 2 A Brief Review of SR theory Stimulus-response theory (or SR theory; see Suppes and Atkinson (1960)) is one of the most successful behavioral learning theories in psychology. Though it has decreased in importance in current psychology, we chose to model SR theory for the following reasons. First, it is based on a rigid trial structure, which permits its concepts to be formally axiomatized, resulting in many important non-trivial but illuminating representation theorems (Suppes, 2002). In fact, the theory is rich enough to represent language in it. Second, despite its few parameters (the learning probability c and the number of stimuli), it has been shown to fit well to empirical data in a variety of experiments. Finally, as we showed in Suppes, de Barros, & Oas (2012), SR theory seems to have natural counterparts at a neuronal level, and is, in some sense still used by neuroscientists (though, sadly, not in its mathematical form). Here we present the mathematical version of SR theory for a continuum of responses, formalized in terms of a stochastic process (we follow Suppes, de Barros, & Oas, 2012). Let (Ω,F , P ) be a probability space, and let Z, S, R, and E be random variables, with Z : Ω → ES S : Ω → S, R : Ω → R, and E : Ω → E, where S is the set of stimuli, R the set of responses, and E the set of reinforcements. Then a trial in SR theory has the following structure: Zn → Sn → Rn → En → Zn+1. (1) The trial structure works the following way. Trial n starts with a certain state of conditioning and a sampled stimulus. Once a stimulus is sampled, a response is computed according to the state of conditioning. Then, reinforcement fol- lows, which can lead (with probability c) to a new state of conditioning for 3 Oscillator model 4 trial n + 1 . In more detail, at the beginning of a trial, the state of condition- 1 i 1 m , . . . , z(n) 1 , its corresponding z(n) i determines on trial n. Once a stimulus Sn = si is sam- m (cid:17). The vector m (cid:17) associates to each stimuli si ∈ S, i = 1, . . . , m, where m = S , . . . , z(n) ing is represented by the random variable Zn = (cid:16)z(n) (cid:16)z(n) is the cardinality of S, a value z(n) pled with probability P (Sn = sisiǫS) = the probability of responses in R by the probability distribution K (cid:16)rz(n) P (cid:16)a1 ≤ Rn ≤ a2Sn = si, Zn,i = z(n) the i-th component of the vector (cid:16)z(n) tion K(cid:16)rz(n) ance and mode z(n) with probability c, i.e. P (cid:16)Zn+1,i = ySn = si, En = y, Zn,i = z(n) P (cid:16)Zn+1,i = z(n) i (cid:17), i.e. i (cid:17) i (cid:17) dx, where k(cid:16)xz(n) k(cid:16)xz(n) m (cid:17). The probability distribu- i (cid:17) is the smearing distribution, and it is determined by its vari- i (cid:17) = c and i (cid:17) = 1 − c. The trial ends with a is the probability density associated to the distribution, and where Zn,i . The next step is the reinforcement En, which is effective Sn = si, En = y, Zn,i = z(n) i (cid:17) = R a2 a1 , . . . , z(n) 1 is i i new (with probability c) state of conditioning Zn+1. 3 Oscillator model In this section we will describe intuitively the oscillator model. We start by arguing for the use of neural oscillators as a way to model the brain at a system level. We then discuss how we can represent in a mathematically sensible way these oscillators. Finally, we show how response computations and learning can be modeled using this theoretical apparatus. Readers interested in more detail are referred to Suppes, de Barros, & Oas (2012). There are many different ways in which researchers try to figure out how the brain works. For example, in cognitive neuroscience, among the most popular research techniques are fMRI (functional magnetic resonance imaging), MEG (magnetoencephalogram), and EEG (electroencephalogram). MEG and EEG measure the electrical activities in the brain, whereas fMRI measures changes in blood flow associated with higher metabolic rates. While fMRI's popularity is due to its better spatial resolution, MEG and EEG present significantly better time resolution. However, what these techniques have in common is that, in order to measure a signal from the brain, they require a large numbers of neurons to fire synchronously. To make our point, let us focus on EEG (though MEG would be adequate too). There are many experiments (see Carvalhaes et al. (2012) and references) showing that the EEG data allow a good representation of language or visual imagery. Thus, neurophysiological evidence points toward language being an activity involving large collections of synchronizing neurons, and we will center our model exactly on this. Before we show how to describe such collections of synchronizing neurons mathematically, it is useful to think about the physical mechanisms of synchro- 3 Oscillator model 5 V (a.u.) TA t0 t0+TA t0+2TA t (a.u.) V (a.u.) TB t0 t0+TB t0+2TAB t (a.u.) Fig. 1: Approximate shape of the action potentials VA and VB as a function of time t for two uncoupled neurons nA and nB firing periodically, with periods TA and TB. For simplicity, we chose a t0 when both neurons fire simultaneously. nization. Let us look first at individual neurons, and then think about ensembles of neurons. Figure 1 shows the qualitative behavior of two neurons nA and nB firing periodically, with TB < TA. What happens if we now couple nA to an excitatory synapse coming from neuron nB? Because t0 + TB < t0 + TA, the excitatory coupling will increase the membrane potential of neuron nA before t0 +TA, causing nA to fire a little earlier than it would if it were not connected to nB. So, excitatory synaptic couplings between neurons can change the timing of coupling, and this timing is changed such that the firings of both neurons approach (in this case, the firing of nA approaches that of nB). In other words, excitatory couplings push nA and nB toward synchronization. In fact, it is pos- sible to prove mathematically that if the number of neurons is large enough, the sum of the many weak synaptic interactions can cause a strong effect, making all neurons fire closer together (Izhikevich, 2007); even when weakly coupled, en- sembles of periodically firing neurons synchronize. It is interesting to note that the argument shown above can be scaled up to distinct collections of neurons. Imagine we have two ensembles of neurons, NA and NB, such that neurons in them synchronize. If neurons in NA and NB become coupled, then the same mechanism as discussed above will be at play, and the ensembles will synchro- nize among themselves. We will come back to this point later, when we talk about response mechanisms. We are now in good shape to introduce the intuition behind the mathemati- cal description for the dynamics of synchronization. One of the main simplifying 3 Oscillator model 6 assumptions we make is that the relevant information coded in the brain is rep- resented by the synchronization of an ensemble of neurons. This ensemble may include tens of thousands of neurons, but because they are synchronized, we can represent them, at least in first approximation, by a single dynamical variable. To understand this, let us think about the simplest case, where an oscillator O (t) can be represented by a sine function1: O (t) = A sin ωt, (2) where ω = ω (t) is its time-dependent frequency. Since ω may be a function of time, the value of O (t) is completely determined by the argument of the sine, i.e. by ϕ = ωt. The quantity ϕ is the phase of the oscillator O (t) = A sin ϕ (t). Since collections of firing neurons have very little variability in its intensity (except, as we see below, when they interfere), we can describe a neural oscillator by its phase. The interaction of a neural oscillator with other neural oscillators may change the evolution of its phase. We emphasize that there is a certain invariance of scale in the above argu- ment: it somehow does not matter how many neurons we have; all that matters is that their amplitude does not vary, that their couplings are strong enough to produce synchronization, and that their dynamics is encoded in the phase. Furthermore, in the same way that individual oscillating neurons synchronize to each other, a collection of coherent neurons can also synchronize to another col- lection of coherent neurons. Since neurons firing coherently may be described approximately by their phase, we can focus on the phase dynamics, instead of being concerned about the full description of the very complex dynamical system. Now, let us look a little more into the details of the mathematics of two synchronizing oscillators. Let us start with two oscillators, O1 (t) and O2 (t), described by their phases ϕ1 and ϕ2. If the two oscillators are uncoupled and their frequency ω is constant, then it is clear from equation (2) that they should satisfy the following set of differential equations, dϕ1 dt dϕ2 dt = ω1, = ω2, (3) (4) where ωi, i = 1, 2, are their natural frequencies. However, if they are weakly coupled, such that their interaction does not affect the overall form of the os- cillations given by O1 (t) and O2 (t) but affects their phase, then equations (3) and (4) need to be modified to include changes to the phase. Furthermore, if the underlying interaction is such that it will make the phases approach each other, such as in the case of synaptically coupled neurons, then it is possible to 1 We use a sine function for simplicity, but the following argument is valid for periodic functions. 3 Oscillator model 7 show that, in first approximation, the modified dynamical equations become dϕ1 dt dϕ2 dt = ω1 − k12 sin (ϕ1 − ϕ2) , = ω2 − k21 sin (ϕ2 − ϕ1) , (5) (6) where kij are the phase coupling strengths. If we extend this to allow for N oscillators, equations (5) and (6) then become dϕi dt = ωi −Xj6=i kij sin (ϕi − ϕj ) . (7) Equation (7) is known as Kuramoto equation (Kuramoto, 1984), and it is widely used to describe complex systems with emergent synchronization. The strength and usefulness of Kuramoto's equation comes from two main points. First, it can be solved under certain symmetric conditions and in the limit of large N , yielding significant insight into the nature of emerging synchronization. Second, a set of weakly-coupled oscillating dynamical systems close to a Andronov-Hopf bifurcation can be described, in first approximation, by Kuramoto-like equations (see Izhikevich (2007)). For our purpose, Kuramoto's equations are a good approximation for the dynamics of coupled neural oscillators. So, we now turn into the discussion of how we can think of stimulus and response as modeled by oscillators, and in particular by Kuramoto's equations. The basic idea is simple. Once a distal stimulus is presented, the perceptual system activates an ensemble of brain neurons, Ns, associated with it. This system itself is described by Kuramoto's equations, and, because it synchronizes, we use its average phase to describe its mean dynamics. If this stimulus elicits a response, the activation of the response neurons via synaptic couplings follows. Responses, as stimuli, are also represented by synchronously firing ensemble of neurons. The selection of a particular response happens when the stimulus oscillator synchronizes in phase with it, and such phase is determined by the relative couplings between stimulus and response oscillators. Let us now look more into its detail. The simplest stimulus-response neural oscillator model requires three os- cillators, Os, Or1 , and Or2 . Os is the oscillator representing firing neurons corresponding to the sampling of a stimulus, and Or1 and Or2 are the response oscillators. Their phases are ϕs, ϕr1 , and ϕr2 . Before we describe their dy- namics, let us go through the process of a response computation. Whenever Os is activated, and subsequently Or1 and Or2 , then the intensity of firings (i.e., the rate of firing, as the individual neuron amplitudes are reasonably stable) in each response oscillator is not only due to its firing, but also to the firings of Os. As we mentioned earlier, a collection of firing neurons may interfere, and in this case, interference means stronger firing rates when in phase, and weaker firing rates when off of phase. Let us analyze this with a mathematically simple 3 Oscillator model example of equal intensity harmonic oscillators, given by Os(t) = A cos (ω0t) = A cos (ϕs(t)) , Or1 (t) = A cos (ω0t + δφ1) = A cos (ϕr1(t)) , Or2 (t) = A cos (ω0t + δφ2) = A cos (ϕr2(t)) . 8 (8) (9) (10) Equations (8) -- (10) represent the case where the oscillators are already syn- chronized with the same frequency ω0 but with relative but constant phase differences δφ1 and δφ2. The mean intensity give us a measure of the excitation carried by the oscillations, and for the superposition of Os(t) and Or1 (t) it is given by I1 = D(Os(t) + Or1 (t))2Et where hf (t)it0 = 1 computation yields t0 = (cid:10)Os(t)2(cid:11)t +(cid:10)Or1 (t)2(cid:11)t + h2Os(t)Or1 (t)it , ∆T R t0+∆T f (t) dt (∆T ≫ 1/ω0) is the time average . A quick I1 = A2 (1 + cos (δφ1)) , and, similarly for I2, I2 = A2 (1 + cos (δφ2)) . Therefore, the intensity depends on the phase difference between the response- computation oscillators and the stimulus oscillator. Now, the maximum intensity of I1 and I2 is 2A2, whereas their minimum intensity is zero. If we think of I1 and I2 as competing possible responses, the maximum difference between them happens when one of their relative phases (with respect to the stimulus oscillator) is zero while the other is π. It is standard to use the contrast, defined by b = I1 − I2 I1 + I2 , (11) as a measure of how different the intensities are. From its definition, b takes values between −1 and 1. When I1 and I2 are as different as possible, b = 1; if, on the other hand, I1 and I2 are the same, b = 0. The contrast provides us with a useful way to think about responses that are between r1 and r2. To see this, let us impose which results in and δφ1 = δφ2 + π ≡ δφ, I1 = A2 (1 + cos (δφ)) , I2 = A2 (1 − cos (δφ)) . (12) (13) (14) In this case, the single parameter δφ is sufficient to determine the contrast, as b = cos (δφ) , (15) 3 Oscillator model 9 0 ≤ δϕ ≤ π. So, the phase difference δφ between stimulus and response oscilla- tors codes a continuum of responses between −1 and 1 (more precisely, because δϕ is a phase, the interval is in the unit circle T, and not in a compact interval in R). For arbitrary intervals (ζ1, ζ2), all that is required is a re-scaling of b. To summarize the above arguments. When a stimulus and response os- cillators activate, they fire periodically,leading to their synchronization with constant phase relation. This phase relation causes interference, which in turn determines the relative strength of the intensities for each response. Thus, re- sponses are determined by the interference of oscillators, which is itself affected by the neural oscillators' couplings. We now examine in more detail the mathematics of the stimulus and response model. Let us look at each step of (1). Sampling When a stimulus sn is sampled, a collection of neurons start firing synchronously, corresponding to the activation of a neural oscillator, Osn . Such activation leads to a spreading of activation to oscillators coupled to the stimulus oscillator, in- cluding the response Or1 and Or2 . Since the selection and activation of Osn involves the perceptual system, we do not attempt to model with neural oscil- lators this step, but simply assume their activation in a way that is consistent with the stochastic process represented in SR theory by the random variable Sn. Furthermore, though it would be important to develop a detailed theory of spreading activation, we do not, as for our current purposes it suffices to simply assume the activation of Or1 and Or2 . Response After the stimulus sn is sampled, the active oscillators evolve for the time in- terval ∆tr, the time it takes to compute a response, according to the following set of Kuramoto differential equations. dϕi dt = ωi −Xi6=j kij sin (ϕi − ϕj + δij) , (16) where kij is the coupling constant between oscillators i and j, and δij is an anti-symmetric matrix representing phase differences, and i and j can be either Osn , Or1 , or Or2 . Here we use the notation where Oi corresponds to a neural oscillator and ϕi to its phase. Equation (16) can be rewritten as dϕi dt = ωi −Xj (cid:2)kE ij sin (ϕi − ϕj) + kI ij cos (ϕi − ϕj)(cid:3) , (17) ij = kij cos (δij) and kI where kE ij = kij sin (δij ), which has an immediate phys- ical interpretation: kE ij corre- sponds to inhibitory ones. These are the 4N excitatory and inhibitory coupling ij corresponds to excitatory couplings, whereas kI 3 Oscillator model strengths between oscillators. dϕi dt = ω0 −Xi6=j (cid:2)kE i,j sin (ϕi − ϕj ) − kI i,j cos (ϕi − ϕj )(cid:3) , 10 (18) where ω0 is their natural frequency. The solutions to (18) and the initial con- ditions randomly distributed at activation give us the phases at time tr,n = ts,n + ∆tr. The coupling strengths between oscillators determine their relative phase locking, which in turn corresponds to the computation of a given response, according to equation (11). Reinforcement and Conditioning As we saw above, the computation of a response depends on the inhibitory and excitatory couplings between neural oscillators. Therefore, when an effective reinforcement Yn corresponding to changes in the conditioning Zn+1 occurs, the coupling strengths change. As with stimulus and responses, we represent a reinforcement by a neural oscillator. Such oscillator, with frequency ωe, is activated during reinforcement, and we assume that it forces the reinforced response-computation and stimulus oscillators to synchronize with the same phase difference of δϕ, while the two response-computation oscillators are kept synchronized with a phase difference of π. Let the reinforcement oscillator be activated on trial n at time te,n, tr,n+1 > te,n > tr,n, for an interval of time ∆te. Let K0 be the coupling strength between the reinforcement oscillator and the stimulus and response-computation oscillators. In order to match the probabilistic SR axiom governing the effectiveness of reinforcement, we also assume that there is a normal probability distribution governing the coupling strength K0 between the reinforcement and the other active oscillators with probability density f (K0) = 1 σK0√2π exp(cid:26)− 1 2σ2 K0 (cid:0)K0 − K 0(cid:1)2(cid:27) . (19) When a reinforcement is effective, all active oscillators phase-reset at te,n, and during reinforcement the phases of the active oscillators evolve according to the following set of differential equations. dϕi dt = ω0 −Xi6=j (cid:2)kE i,j sin (ϕi − ϕj) − kI i,j cos (ϕi − ϕj)(cid:3) −K0 sin (ϕi − ωet + Φi) , (20) where Φsn − Φr1 = δϕ and Φr1 − Φr2 = π. The excitatory couplings are rein- forced if the oscillators are in phase with each other, according to the following equations. dkE i,j dt = ǫ (K0)(cid:2)α cos (ϕi − ϕj) − kE i,j(cid:3) . (21) 4 Final remarks 11 Similarly, for inhibitory connections, if two oscillators are perfectly off sync, then we have a reinforcement of the inhibitory connections. dkI i,j dt = ǫ (K0)(cid:2)α sin (ϕi − ϕj) − kI i,j(cid:3) , In the above equations, ǫ (K0) = (cid:26) 0 if K0 < K ′ ǫ0 otherwise, (22) (23) where ǫ0 ≪ ω0, α and K0 are constant during ∆te, and K ′ is a threshold constant throughout all trials. We can think of K ′ as a threshold below which the reinforcement oscillator has no (or very little) effect on the stimulus and response-computation oscillators. For large enough values of ∆te, the behavioral probability parameter c of effective reinforcement mentioned above is, from (19) and (23), reflected in the equation: c = Z ∞ K ′ f (K0) dK0. (24) This relationship comes from the fact that, if K0 < K ′, there is no effective learning from reinforcement, since there are no changes to the couplings due to (21) -- (22), and (18) describing the oscillators' behavior. Intuitively K ′ is the effectiveness parameter: the larger it is, the smaller the probability of effective reinforcement. 4 Final remarks In this paper we described the neural oscillator model presented in Suppes, de Barros, & Oas (2012), with particular emphasis to the physics and intuition behind many of the processes represented by equations (18). To summarize it, the coded phase differences were used to model a continuum of responses within SR theory in the following way. At the beginning of a trial a stimulus oscillator is activated, and with it the response oscillators. Then, the coupled oscillator system evolves according to (18) if no reinforcement is present, and according to (20) -- (22) if reinforcement is present. The coupling constants and the conditioning of stimuli are not reset at the beginning of each trial, and changes to couplings correspond to effective reinforcement. Because of the finite amount of time for a response, the probabilistic characteristics of the initial conditions lead to the smearing of the phase differences after a certain time, with an effect similar to that of the smearing distribution in the SR model for a continuum of responses (Suppes, 1959). We emphasize that in this paper we focused mainly on the physical basis of our model, and did not go much into mathematical detail. Furthermore, in Suppes, de Barros, & Oas (2012) we applied the neural oscillator model to many different experimental situations illustrated in the literature, whereas here we 4 Final remarks 12 did not address in detail any empirical data. Interested readers are referred to our original paper. SR theory has enjoyed tremendous success in the past, and, in a certain sense, its main features are still present in modern day neuroscience. We be- lieve that by showing how neurons may result in theoretical structures that are somewhat similar to SR ones, as done in Suppes, de Barros, & Oas (2012), we can provide the basis for an extension of SR theory that could be considered more realistic. For example, in our model, many parameters, such as time of response, frequency of oscillations, coupling strengths, etc., were fixed based on reasonable assumptions. However, a more detailed and systematic study should be able to relate such parameters to either underlying physiological constraints or to behavioral variations, thus opening up the possibilities for new empiri- cal studies that go beyond SR theory. Also, in our model we postulated many features without showing or proving their dynamics from underlying neuronal dynamics. This was the case for the activation of a stimulus and the spreading of activation of a stimulus and responses. A more detailed theory based on neural oscillators of such dynamics would certainly provide interesting empirical tests. Finally, the use of neural oscillators and interference may also help explain certain aspects of cognition that are considered "non-classical." The distinction between classical and quantum behavior is a subtle one, and still not yet un- derstood. For example, a well studied quantum-like decision making process is the violation of Savage's sure-thing principle, shown in a series of experiments by Tversky and Shaffir (Shafir and Tversky, 1992; Tversky and Shafir, 1992). Similar violations do not need any quantum-like representation in the form of a Hilbert space, as proposed in the literature, but instead can be obtained by interference of neural oscillators (de Barros, 2012b). Furthermore, the use of neural oscillator interference even leads to predictions that are not compatible with a Hilbert space structure (de Barros, 2012a), suggesting that the use of quantum-like processes is not as quantum as many would wish. References Adolphs, R. (2003). Cognitive neuroscience of human social behaviour. Nature Reviews Neuroscience, 4(3):165 -- 178. Carvalhaes, C. G., de Barros, J. A., Perrau-Guimaraes, M., and Suppes, P. (2012). Using the scalp electric field to improve recognition of brain processes. In preparation. de Barros, J. A. (2012a). Joint probabilities and quantum cognition. In Khren- nikov, A., editor, Proceedings of Quantum Theory: Reconsiderations of Foun- dations - 6, Växjö, Sweden. Institute of Physics. To appear. de Barros, J. A. (2012b). Quantum-like model of behavioral response computa- tion using neural oscillators. arXiv:1207.0033. Submitted to BioSystems. 4 Final remarks 13 de Barros, J. A. and Suppes, P. (2009). Quantum mechanics, interference, and the brain. Journal of Mathematical Psychology, 53(5):306 -- 313. Izhikevich, E. M. (2007). Dynamical Systems in Neuroscience: The Geometry of Excitability and Bursting. The MIT Press, Cambridge, Massachusetts. Kandel, E. R., Schwartz, J. H., and Jessell, T. M. (2000). Principles of neural science. McGraw-Hill, New York, NY, 4th edition. Kuramoto, Y. (1984). Chemical Oscillations, Waves, and Turbulence. Dover Publications, Inc., Mineola, New York. Shafir, E. and Tversky, A. (1992). Thinking through uncertainty: Nonconse- quential reasoning and choice. Cognitive Psychology, 24(4):449 -- 474. Suppes, P. (1959). Stimulus sampling theory for a continuum of responses. In Arrow, K., Karlin, S., and Suppes, P., editors, Mathematical Methods in the Social Sciences, pages 348 -- 365. Stanford University Press, Stanford, California. Suppes, P. (2002). Representation and Invariance of Scientific Structures. CSLI Publications, Stanford, California. Suppes, P. and Atkinson, R. C. (1960). Markov learning models for multiperson interactions. Stanford University Press, Stanford, California. Suppes, P. and de Barros, J. (2007). Quantum mechanics and the brain. Quan- tum Interaction: Papers from the AAAI Spring Symposium, Technical Report SS-07, 8:75 -- 82. Suppes, P., de Barros, J. A., and Oas, G. (2012). Phase-oscillator computations as neural models of stimulus -- response conditioning and response selection. Journal of Mathematical Psychology, 56(2):95 -- 117. Tversky, A. and Shafir, E. (1992). The disjunction effect in choice under uncer- tainty. Psychological Science, 3(5):305 -- 309. Vassilieva, E., Pinto, G., de Barros, J., and Suppes, P. (2011). Learning pat- tern recognition through quasi-synchronization of phase oscillators. Neural Networks, IEEE Transactions on, 22(1):84 -- 95.
1102.0166
1
1102
2011-02-01T14:32:38
Hebbian learning of recurrent connections: a geometrical perspective
[ "q-bio.NC", "nlin.AO" ]
We show how a Hopfield network with modifiable recurrent connections undergoing slow Hebbian learning can extract the underlying geometry of an input space. First, we use a slow/fast analysis to derive an averaged system whose dynamics derives from an energy function and therefore always converges to equilibrium points. The equilibria reflect the correlation structure of the inputs, a global object extracted through local recurrent interactions only. Second, we use numerical methods to illustrate how learning extracts the hidden geometrical structure of the inputs. Indeed, multidimensional scaling methods make it possible to project the final connectivity matrix on to a distance matrix in a high-dimensional space, with the neurons labelled by spatial position within this space. The resulting network structure turns out to be roughly convolutional. The residual of the projection defines the non-convolutional part of the connectivity which is minimized in the process. Finally, we show how restricting the dimension of the space where the neurons live gives rise to patterns similar to cortical maps. We motivate this using an energy efficiency argument based on wire length minimization. Finally, we show how this approach leads to the emergence of ocular dominance or orientation columns in primary visual cortex. In addition, we establish that the non-convolutional (or long-range) connectivity is patchy, and is co-aligned in the case of orientation learning.
q-bio.NC
q-bio
Hebbian learning of recurrent connections: a geometrical perspective Mathieu N. Galtier ∗ Olivier D. Faugeras † Paul C. Bressloff ‡ May 29, 2018 Abstract: We show how a Hopfield network with modifiable recurrent connec- tions undergoing slow Hebbian learning can extract the underlying geometry of an input space. First, we use a slow/fast analysis to derive an averaged system whose dynamics derives from an energy function and therefore always converges to equilibrium points. The equilibria reflect the correlation structure of the in- puts, a global object extracted through local recurrent interactions only. Second, we use numerical methods to illustrate how learning extracts the hidden geomet- rical structure of the inputs. Indeed, multidimensional scaling methods make it possible to project the final connectivity matrix on to a distance matrix in a high- dimensional space, with the neurons labelled by spatial position within this space. The resulting network structure turns out to be roughly convolutional. The resid- ual of the projection defines the non-convolutional part of the connectivity which is minimized in the process. Finally, we show how restricting the dimension of the space where the neurons live gives rise to patterns similar to cortical maps. We motivate this using an energy efficiency argument based on wire length min- imization. Finally, we show how this approach leads to the emergence of ocular ∗Corresponding author: [email protected]. NeuroMathComp Project Team, INRIA Sophia-Antipolis M´editerran´ee, 2004 route des Lucioles-BP 93, 06902 Sophia Antipolis, France †NeuroMathComp Project Team, INRIA Sophia-Antipolis M´editerran´ee, 2004 route des Lucioles-BP 93, 06902 Sophia Antipolis, France ‡Department of Mathematics, University of Utah, 155 South 1400 East, Salt Lake City, Utah 84112, USA. Mathematical Institute, University of Oxford, 24-29 St. Giles', Oxford OX1 3LB, UK 1 dominance or orientation columns in primary visual cortex. In addition, we es- tablish that the non-convolutional (or long-range) connectivity is patchy, and is co-aligned in the case of orientation learning. Keywords: averaging, energy minimization, multidimensional scaling, cortical maps correlation-based Hebbian learning, Hopfield networks, temporal 1 Introduction Activity-dependent synaptic plasticity is generally thought to be the basic cellular substrate underlying learning and memory in the brain. Donald Hebb [Hebb, 1949] postulated that learning is based on the correlated activity of synaptically con- nected neurons: if both neurons A and B are active at the same time, then the synapses from A to B and B to A should be strengthened proportionally to the product of the activity of A and B. However, as it stands, Hebb's learning rule diverges. Therefore, various modification of Hebb's rule have been developed, which basically take one of three forms (see [Gerstner and Kistler, 2002] and [Dayan and Abbott, 2001]): first, a decay term can be added to the learning rule so that each synaptic weight is able to "forget" what it previously learned. Sec- ond, each synaptic modification can be normalized or projected on different sub- spaces. These constraint–based rules may be interpreted as implementing some form of competition for energy between dendrites and axons, see [Miller, 1996, Miller and MacKay, 1996] and [Ooyen, 2001] for details. Third, a sliding thresh- old mechanism can be added to Hebbian learning. For instance, a post-synaptic threshold rule consists in multiplying the presynaptic activity and the subtraction of the average postsynaptic activity from its current value, which is referred as covariance learning ([Sejnowski and Tesauro, 1989]). Probably the best known of these rules is the BCM rule [Bienenstock et al., 1982]. It should be noted that history-based rules can also be defined without changing the qualitative dy- namics of the system: instead of considering the instantaneous value of the neu- rons' activity, these rules consider its weighted mean over a time window (see [Foldi´ak, 1991, Wallis and Baddeley, 1997]). Recent experimental evidence sug- gests that learning may also depend upon the precise timing of action potentials [Bi and Poo, 2001]. Contrary to most Hebbian rules that only detect correlations, these rules can also encode causal relationships in the patterns of neural activa- tion. However, the mathematical treatment of these spike timing dependent rules is much more difficult than rate based ones. 2 Hebbian-like learning rules have often been studied within the framework of unsupervised feedfoward neural networks [Oja, 1982, Bienenstock et al., 1982, Miller and MacKay, 1996, Dayan and Abbott, 2001]. They also form the basis of most weight-based models of cortical development, assuming fixed lateral con- nectivity (e.g. mexican hat) and modifiable vertical connections (see the review of [Swindale, 1996])1. In these developmental models, the statistical structure of input correlations provides a mechanism for spontaneously breaking some un- derlying symmetry of the neuronal receptive fields leading to the emergence of feature selectivity. When such correlations are combined with fixed intracorti- cal interactions, there is a simultaneous breaking of translation symmetry across cortex leading to the formation of a spatially periodic cortical feature map. A re- lated mathematical formulation of cortical map formation has been developed in [Takeuchi and Amari, 1979, Bressloff, 2005] using the theory of self–organizing neural fields. Although very irregular, the two-dimensional cortical maps ob- served at a given stage of development, can be unfolded in higher dimensions to get smoother geometrical structures. Indeed, [Bressloff et al., 2001] suggested that the network of orientation pinwheels in V1 is a direct product between a circle for orientation preference and a plane for position, based on a modification of the icecube model of Hubel and Wiesel [Hubel and Wiesel, 1977]. From a more ab- stract geometrical perspective, Petitot [Petitot, 2003] has associated such a struc- ture to a 1-jet space and used this to develop some applications to computer vision. More recently, [Bressloff and Cowan, 2003] and [Chossat and Faugeras, 2009] have considered more complex geometrical structures such as spheres and hyperbolic surfaces that incorporate additional stimulus features such as spatial frequency and textures, respectively. In this paper, we show how geometrical structures related to the distribution of inputs can emerge through unsupervised Hebbian learning applied to recur- rent connections in a rate-based Hopfield network. Throughout this paper, the inputs are presented as an external non-autonomous forcing to the system and not an initial condition as is often the case in Hopfield networks. It has previously been shown that, in the case of a single fixed input, there exists an energy func- tion that describes the joint gradient dynamics of the activity and weight variables [Dong and Hopfield, 1992]. This implies that the system converges to an equilib- rium during learning. We use averaging theory to generalize the above result to the case of multiple inputs, under the adiabatic assumption that Hebbian learning 1There have only been a few computational studies that consider the joint development of lateral and vertical connections [Bartsch and Van Hemmen, 2001, Miikkulainen et al., 2005]. 3 occurs on a much slower time scale than both the activity dynamics and the sam- pling of the input distribution. We then show that the equilibrium distribution of weights, when embedded into Rk for sufficiently large integer k, encodes the geo- metrical structure of the inputs. Finally, we numerically show that the embedding of the weights in two dimensions (k = 2) gives rise to patterns that are quali- tatively similar to experimentally observed cortical maps, with the emergence of features columns and patchy connectivity. Although the mathematical formalism we introduce here could be extended to most of the rate-based Hebbian rules in the literature, we present the theory for Hebbian learning with decay because of the simplicity of the resulting dynamics. Note that the use of geometrical objects to describe the emergence of connec- tivity patterns has previously been put forward by Amari in a different context. Based on the theory of information geometry, Amari considers the geometry of the set of all the networks and defines learning as a trajectory on this manifold for perceptron networks in the framework of supervised learning [Amari, 1998] or for unsupervised Boltzmann Machines [Amari et al., 1992]. He uses differential and Riemannian geometry to describe an object which is at a larger scale than the cortical maps this paper is focusing on. Moreover, Zucker and colleagues are currently developing a non-linear dimen- sionality reduction approach to caracterize the statistics of natural visual stimuli (see [Lawlor and Zucker, 2010, Coifman et al., 2005]). Although they do not use learning neural networks and stay closer to the field of computer vision than this paper, it turns out their approach is similar to the geometrical embedding approach we are using. The structure of the paper is as follows. In section 2, we formally introduce the model. We derive the averaged system in section 3, which then allows us to study the stability of the learning dynamics in the presence of multiple inputs by constructing an appropriate energy function. We adress stability in section 4. In section 5 we determine the geometrical structure of the equilibrium weight distribution and show how it reflects the structure of the inputs. We also relate this approach to the emergence of cortical maps. Finally, the results are discussed in section 6. 4 2 Model 2.1 Neural network evolution A neural mass corresponds to a mesoscopic coherent group of neurons. It is con- venient to consider them as building blocks, first for computational simplicity, second for their direct relationship to macroscopic measurements of the brain (EEG, MEG and Optical imaging) which average over numerous neurons, and third because one can functionally define coherent groups of neurons within cor- tical columns. For each neural mass i ∈ {1..N}, define the mean membrane potential Vi(t) at time t. The instantaneous population firing rate νi(t) is linked to the membrane potential through the relation νi(t) = s(cid:0)Vi(t)(cid:1), where s is a smooth sigmoid function. In the following, we choose s(v) = Sm 1 + exp(cid:0) − 4S′ m(v − φ)(cid:1) , where Sm, S′ and the offset of the sigmoid. m and φ are respectively the maximal firing rate, the maximal slope Consider a Hopfield network of neural masses described by the equation (1) (2) dVi dt (t) = −αVi(t) + NXj=1 Wij(t) s(cid:0)Vj(t)(cid:1) + Ii(t). The first term roughly corresponds to the intrinsic dynamics of the neural mass: it decays exponentially to zero at a rate α if it receives neither external inputs nor spikes from the other neural masses. We will fix the units of time by setting α = 1. The second term corresponds to the rest of the network sending information through spikes to the given neural mass i, with Wij(t) the effective synaptic weight from neural mass j. The synaptic weights are time–dependent because they evolve according to a continuous time Hebbian learning rule (see below). The third term Ii(t) corresponds to an external input to neural mass i, e.g. information extracted by the retina or thalamo-cortical connections. We take the inputs to be piecewise constant in time, that is, at regular time intervals a new input is presented to the network. In this paper, we will assume that the inputs are chosen by peridodically cycling through a given set of M inputs. An alternative approach would be to randomly select each input from a given probability distribution [Geman, 1979]. It is convenient to introduce vector notation by representing the time–dependent membrane potentials by V ∈ C 1(R+, RN ), the time–dependent external inputs 5 by I ∈ C 0(R+, RN ), and the time–dependent network weight matrix by W ∈ C 1(R+, RN ×N ). We can then rewrite the above system of ordinary differential equations as a single vector-valued equation dV dt = −V + W · S(V ) + I, (3) where S : RN → RN corrresponds to the term by term application of the sigmoid S, i.e. S(V )i = s(Vi). 2.2 Correlation-based Hebbian learning The synaptic weights are assumed to evolve according to a correlation–based Heb- bian learning rule of the form dWij dt = ǫ(s(Vi)s(Vj) − µWij), (4) where ǫ is the learning rate, and we have included a decay term in order to stop the weights from diverging. In order to rewrite the above equation in a more compact vector form, we introduce the tensor (or Kronecker) product S(V ) ⊗ S(V ) so that in component form [S(V ) ⊗ S(V )]ij = S(V )iS(V )j, (5) where S is treated as a mapping from RN to RN . The tensor product implements Hebb's rule that synaptic modifications involve the product of postynaptic and presynaptic firing rates. We can then rewrite the combined voltage and weight dynamics as the following non–autonomous (due to time–dependent inputs) dy- namical system: Σ :  dV dt dW dt = −V + W · S(V ) + I = ǫ(cid:16)S(V ) ⊗ S(V ) − µW(cid:17). (6) Let us make few remarks about the existence and uniqueness of solutions. First, boundedness of S implies boundedness of the system Σ. More precisely, if I is bounded, then the solutions are bounded. To prove this, note that the right hand side of the equation for W is the sum of a bounded term and a linear decay 6 term in W . Therefore, W is bounded and hence the term W ·S(V ) is also bounded. The same reasoning applies to V . S being Lipschitz continuous implies that the right hand side of the system is Lipschitz. This is sufficient to prove existence and uniqueness of the solution by applying the Cauchy-Lipschitz theorem. In the following, we will derive an averaged autonomous dynamical system Σ′, which will be well-defined for the same reasons. 3 Averaging the system We will show that system Σ can be approximated by an autonomous Cauchy prob- lem which will be much more convenient to handle. This averaging method makes the most of multiple time–scales in the system. First, it is natural to consider that learning occurs on a much slower time–scale than the evolution of the membrane potentials (as determined by α), i.e. ǫ ≪ 1. (7) Second, an additional time-scale arises from the rate at which the inputs are sam- pled by the network. That is, the network cycles periodically through M fixed inputs, with the period of cycling given by T . It follows that I is T –periodic, piecewise constant. We assume that the sampling rate is also much slower than the evolution of the membrane potentials, M T ≪ 1. (8) Finally, we assume that the period T is small compared to the time-scale of the learning dynamics, ǫ ≪ . (9) 1 T We can now simplify the system Σ by applying Tikhonov's theorem for slow/fast systems, and then classical averaging methods for periodic systems. 3.1 Tikhonov's theorem Tikhonov's theorem ([Tikhonov, 1952] and [Verhulst, 2007] for a clear introduc- tion) deals with slow/fast systems. It says the following: 7 Theorem 3.1. Consider the initial value problem x = f (x, y, t), x(0) = x0, x ∈ Rn, t ∈ R+ ǫ y = g(x, y, t), y(0) = y0, y ∈ Rm Assume that: 1. A unique solution of the initial value problem exists and we suppose, this holds also for the reduced problem x = f (x, y, t), x(0) = x0 0 = g(x, y, t) with solutions ¯x(t), ¯y(t). 2. The equation 0 = g(x, y, t) is solved by ¯y(t) = φ(x, t), where φ(x, t) is a continuous function and an isolated root. Also suppose that ¯y(t) = φ(x, t) is an asymptotically stable solution of the equation dy dτ = g(x, y, τ ) that is uniform in the parameters x ∈ Rn and t ∈ R+. 3. y(0) is contained in an interior subset of the domain of attraction of ¯y. Then we have limǫ→0 xǫ(t) = ¯x(t), 0 ≤ t ≤ L limǫ→0 yǫ(t) = ¯y(t), 0 ≤ d ≤ t ≤ L with d and L constants independent of ǫ. In order to apply Tikhonov's theorem directly to the system Σ, we first need to rescale time according to t → ǫt. This gives ǫ dV dt dW dt = −V + W · S(V ) + I = S(V ) ⊗ S(V ) − µW. Tikhonov's theorem then implies that solutions of Σ are close to solutions of the reduced system (in the unscaled time variable) ( V (t) = W · S(cid:0)V (t)(cid:1) + I(t) W = ǫ(cid:16)S(V ) ⊗ S(V ) − µW(cid:17), 8 (10) provided that the dynamical systems Σ in equation (6), and equation (10) are well defined. It is easy to show that both systems are Lipschitz because of the properties of S. Following [Faugeras et al., 2008], we know that if S′ mkW k < 1, (11) then there exists an isolated root ¯V : R+ → RN of the equation V = W · S(V ) + I and ¯V is asymptotically stable. Equation (11) corresponds to the weakly con- nected case. Moreover, the initial condition belongs to the basin of attraction of this single fixed point. Note that we require M T ≪ 1 so that the membrane po- tentials have sufficient time to approach the equilibrium associated with a given input before the next input is presented to the network. In fact, this assumption make it reasonable to neglect the transient activity dynamics due to the switching between inputs. 3.2 Periodic averaging The system given by equation (10) corresponds to a differential equation for W with T -periodic forcing due to the presence of V on the right–hand side. Since T ≪ ǫ−1, we can use classical averaging methods to show that solutions of (10) are close to solutions of the following autonomous system on the time-interval [0, 1 ǫ ] (which we suppose large because ǫ << 1) V (t) = W · S(V (t)) + I(t) Σ0 :  dW dt T Z T = ǫ(cid:16) 1 0 S(V (s)) ⊗ S(V (s))ds − µW (t)(cid:17). (12) It follows that solutions of Σ are also close to solutions of Σ0. Finding the explicit solution V (t) for each input I(t) is difficult and requires fixed points methods, e.g. a Picard algorithm. Therefore, we will consider yet another system Σ′ whose solutions are also close to Σ0 and hence Σ. In order to construct Σ′ we need to introduce some additional notation. Let us label the M inputs by I (a), a = 1, . . . , M and denote by V (a) the fixed point solution of the equation V (a) = W · S(V (a)) + I (a). Given the periodic 9 sampling of the inputs, we can rewrite (12) as V (a) = W · S(V (a)) + I (a) dW dt = ǫ(cid:16) 1 M MXa=1 S(V (a)) ⊗ S(V (a)) − µW (t)(cid:17). (13) If we now introduce the N × M matrices V and I with components Via = V (a) and Iia = I (a) , then we can eliminate the tensor product and simply write (13) in the matrix form i i V = W · S(V) + I dW dt = ǫ(cid:16) 1 M S(V) · S(V)T − µW (t)(cid:17), where S(V) ∈ RN ×M such that [S(V)]ia = s(V (a) ). A second application of Tikhonov's theorem (in the reverse direction) then establishes that solutions of the system Σ0 (written in the matrix form (14)) are close to solutions of the matrix system i (14) (15) Σ′ :  dV dt dW dt = −V + W · S(cid:0)V(cid:1) + I = ǫ(cid:16) 1 S(V) · S(V)T − µW (t)(cid:17) M m µ . Therefore, equation (11) says that if S2 In the remainder of the paper we will focus on the system Σ′ whose solutions are close to those of the original system Σ provided condition (11) is satisfied, i.e. the network is weakly connected. Clearly, the fixed points (V ∗, W ∗) of sys- tem Σ satisfy kW ∗k ≤ S2 mS ′ µ < 1 then m Tikhonov's theorem can be applied and systems Σ and Σ′ can be reasonably con- sidered as good approximations of each other. The advantage of the averaged sys- tem Σ′ is that is given by autonomous ordinary differential equations. Moreover, since it is Lipschitz continuous, it leads to a well-posed Cauchy problem. Finally, note that it is straighforward to extend our approach to time-functional rules (e.g. sliding threshold or BCM rules as described in [Bienenstock et al., 1982]) which, in this new framework, would be approximated by simple ordinary differential equations (as opposed to time-functional differential equations) provided S is re- defined appropriately. 10 1 5 10 25 50 100 250 error in % 105 90 75 60 45 30 15 0 Figure 1: Percentage of error between final connectivities for the exact and aver- aged system. 3.3 Simulations To illustrate the above approximation, we simulate a simple network with both exact, i.e. Σ, and averaged ,i.e. Σ′, evolution equations. For these simulations, the network consists of N = 10 fully-connected neurons and is presented with M = 10 different random inputs taken uniformly in the intervals [0, 1]N. For this simulation we use s(x) = 1+e−4(x−1) , and µ = 10. Figure 1. shows the percentage of error between final connectivities for different values of ǫ and T /M. Figure 2 shows the temporal evolution of the norm of the connectivity for both the exact and averaged system for T = 103 and ǫ = 10−3. 1 4 Stability 4.1 Liapunov function In the case of a single fixed input (M = 1), the systems Σ and Σ′ are equivalent and reduce to the neural network with adapting synapses previously analyzed by [Dong and Hopfield, 1992]. Under the additional constraint that the weights are symmetric (Wij = Wji), these authors showed that the simultaneous evolution of the neuronal activity variables and the synaptic weights can be re-expressed as a 11 0.040 0.035 0.030 0.025 0.020 0.015 0.010 0.005 averaged system exact system 0 5000 10000 15000 20000 Figure 2: Temporal evolution of the norm of the connectivities of the exact system Σ and averaged system Σ′. gradient dynamical system that minimizes a Liapunov or energy function of state. We can generalize their analysis to the case of multiple inputs (M > 1) and non- symmetric weights using the averaged system Σ′. That is, following along similar lines to [Dong and Hopfield, 1992], we introduce the energy function E(U, W ) = − 1 2 where U = S(V), kW k2 = hW, W i =Pi,j W 2 ij, hU, W · Ui = hU, W · Ui − hI, Ui + h1, S−1(cid:0)U(cid:1)i + MXa=1 MXa=1 h1, S−1(cid:0)U(cid:1)i = NXi=1Z U (a) U (a) i WijU (a) NXi=1 MXa=1 S−1(ξ)dξ. , j hI, Ui = i 0 Mµ 2 kW k2 (16) I (a) i U (a) i (17) NXi=1 (18) and In contrast to [Dong and Hopfield, 1992], we do not require a priori that the weight matrix is symmetric. However, it can be shown that the system always converges to a symmetric connectivity pattern. More precisely, A =n(V, W ) ∈ RN ×M × RN ×N : W = W To is an attractor of the system Σ′. A proof can be found in appendix 8.1. It can then be shown that on A (symmetric weights), E is a Liapunov function of the dynamical system Σ′, that is, dE dt ≤ 0, and dE dt = 0 =⇒ dY dt = 0, Y = (V, W )T . 12 The boundedness of E and the Krasovskii-LaSalle invariance principle then im- plies that the system converges to an equilibrium [Khalil and Grizzle, 1996]. We thus have Theorem 4.1. The initial value problem for the system Σ′ with Y(0) ∈ H, con- verges to an equilibrium state. Proof. See appendix 8.2 It follows that neither oscillatory nor chaotic attractor dynamics can occur. 4.2 Linear stability Although we have shown that there are stable fixed points, not all of the fixed points are stable. However, we can apply a linear stability analysis on the system Σ′ to derive a simple sufficient condition for a fixed point to be stable. The method we use in the proof could be extended to more complex rules. The proof reveals the significant role played by the Kronecker product in Hebbian learning. Theorem 4.2. The equilibria of system Σ′ satisfy: (cid:26) V ∗ = 1 W ∗ = 1 µM S(V ∗) · S(V ∗)T · S(V ∗) + I µM S(V ∗) · S(V ∗)T and a sufficient condition for stability is 3S′ mkW ∗k < 1 provided 1 > ǫµ which is most probably the case since ǫ << 1. Proof. See appendix 8.3 (19) (20) This condition is strikingly similar to that derived in [Faugeras et al., 2008]. In fact, condition (20) is stronger than the contracting condition (11). It says the network may converge to a weakly connected situation. It justifies the averag- ing method by saying that we remain in the domain of validity of the averaging method. It also says that the dynamics of V is likely (because the condition is only sufficient) to be contracting and therefore subject to no bifurcations: a fully recurrent learning neural network is likely to have a "simple" dynamics. 13 5 Geometrical structure of equilibrium points 5.1 Learning the correlation matrix of the inputs It follows from equation (19) that the equilibrium weight matrix W ∗ is given by the correlation matrix of the firing rates. Moreover, in the case of sufficiently large inputs, the matrix of equilibrium membrane potentials satisfies V ∗ ≈ I. More precisely, if S(I (a) for all a = 1, . . . , M and i = 1, . . . , N, then we can generate an iterative solution for V ∗ of the form ) ≪ I (a) i i V ∗ = I + 1 µ S(cid:0)I(cid:1) · S(cid:0)I(cid:1)T · S(cid:0)I(cid:1) + h.o.t. On the other hand, if the inputs are comparable in size to the synaptic weights, then there is no explicit solution for V ∗. Roughly speaking, we observe that the connection term has the role of "smoothing" the solution. Therefore, if a Gaussian is presented to the network (as the only input), the membrane potential is likely to be another Gaussian with a larger variance. If no input is presented to the network (I = 0), then S(0) 6= 0 implies that the activity is non-zero, that is, there is spontaneous activity. Combining these observations, we see that the network roughly extracts and stores the correlation matrix of the strongest inputs within the weights of the network. 5.2 From a symmetric connectivity matrix to a convolutional network So far neurons have been identified by a label i ∈ {1..N}; there is no notion of geometry or space in the preceding results. However, as we show below, the inputs may contain a spatial structure that can be encoded by the connectivity. In this section, we propose a mechanism to unveil the hidden geometrical structure within the connectivity. More specifically, we want to find an integer k ∈ N and N points in Rk, denoted by xi, i ∈ {1, . . . , N}, so that the connectivity can roughly be written as W ∗ ij ≃ exp(−kxi − xjk2). In other words, we interpret the final connectivity as a matrix describing the distance between the neurons living in a k-dimensional space. However, W ∗ is not always a distance matrix, therefore, it is natural to project it on the set of distance matrices. Finding the best fit of W ∗ to a distance matrix is usually called multidimensional scaling. This set of methods is reviewed in [Borg and Groenen, 2005]. 14 max the max. Second, define a bijective kernel function on x ∈ R+ such that largest component of W ∗: W ∗ Wii/W ∗ K(x) = W ∗ First, definecW ∈ RN ×N as W ∗ whose diagonal terms are set to W ∗ ij = NijcWij with Nij = 1 if i 6= j and Nii = maxe−x/σ2. Given thatcW is non-negative, we define the matrix D = K −1(cW ) corresponding to the application of the inverse of K to each component ofcW . As said before, we want to find k ∈ N and xi ∈ Rk for i ∈ {1, . . . , N} However, we can compute the projection of the symmetric matrixcW onto the set of distance matrices by applying multidimensional scaling methods as described in [Borg and Groenen, 2005]. We use the stress majorization or SMACOF algo- rithm for the stress1 cost function throughout the article. In other words, we can find the distance matrix D⊥ such that kDqk = kD − D⊥k is minimal. Therefore, so that D is a distance matrix, Dij = kxi − xjk2. In general, this is not possible. cWij = K(Dq)ijK(D⊥)ij. Define M such that M(xi, xj) = K(Dq)ij Nij and Gσ a Gaussian with a standard deviation equal to σ. In spatial coordinates W ∗(xi, xj) = M(xi, xj) Gσ(kxi − xjk) (21) Multidimensional scaling methods consist in minimizing the contribution of M in the preceding equation. Hence, we refer to M as the non-convolutional connec- tivity. A position xi ∈ Rk is associated to each neuron i ∈ {1, . . . , N} such that (cid:0)W · S(V )(cid:1)i = W ∗ ijS(Vj) = NXj=1 NXj=1 M(xi, xj) Gσ(kxi − xjk) S(V (xj)) (22) In particular, we can assume that k is large enough for kDqk to be very small. Moreover, if the neurons are equally excited on average (i.e. the diagonal of W ∗ is already equal to W ∗ maxId), then it is reasonable to consider that M(xi, xj) = 1 leading to the following convolutional product W · S(V ) = Gσ(k.k) ∗ S(V ) Therefore, in the space defined by the xi ∈ Rk the connectivity is close to being convolutional. 5.3 Unveiling the geometrical structure of the inputs We hypothesize that the space defined by the xi reflects the underlying geometrical structure of the inputs. We have not found a way to prove this so we only provide 15 numerical examples that illustrate this claim. In the following examples, we relate the geometry of the manifold suggested by the xi to the network inputs. Thus we feed the network with inputs having a defined geometrical structure and then show how this structure can be extracted from the connectivity by the method above. However, it is by extracting the structure from unknown inputs that these networks might reveal themselves useful. Therefore, the following is only a (numerical) proof of concept. We assume the inputs to be uniformly distributed over a manifold Ω with fixed geometry. This strong assumption amounts to considering that the feedforward connectivity (which we do not consider here) has already properly filtered the information coming from the sensory organs. More precisely, define the set of inputs by the matrix I ∈ RN ×M such that I (a) i = f (kyi − zakΩ) where the za are uniformly distributed points over Ω, the yi are the positions on Ω that "label" the ith neuron, and f is a decreasing function on R+. The norm k.kΩ is the natural norm defined over the manifold Ω. For simplicity, assume f (x) = fσ(x) = Ae− x2 so that the inputs are localized bell-shaped bumps on the shape Ω. σ2 5.3.1 Planar retinotopy 0 0 20 40 60 80 20 40 60 80 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0 0 20 40 60 80 20 40 60 80 0.0032 0.0028 0.0024 0.0020 0.0016 0.0012 0.0008 0.0004 Figure 3: Plot of planar retinotopic inputs on Ω = [0, 1]×[0, 1] (left) and final con- nectivity matrix of the system Σ′ (right). The parameters used for this simulation are s(x) = 1+e−4(x−1) , l = 1, µ = 10, ǫ = 0.001, N = M = 100, σ = 4. 1 16 We consider a set of Gaussian inputs uniformly distributed over a two-dimensional plane, e.g. Ω = [0, 1] × [0, 1]. For simplicity, we take N = M = K 2 and set za = yi for i = a, a ∈ {1, . . . , M}. (The numerical results show an identical structure for the final connectivity when the yj correspond to random points, but the analysis is harder). In the simpler case of one-dimensional Gaussians with N = M = K, the input matrix takes the form I = Tfσ, where Tf is a symmetric Toeplitz matrix: f (0) f (1) f (2) ... f (1) f (0) f (1) ... f (2) f (1) f (0) ... f (K) f (K − 1) f (K − 2) · · · f (2) f (1) ... · · · · · · · · · · · · ... · · · f (K) f (K − 1) f (K − 2) ... f (0) (23)  Tf =  It follows that I (a) In the two-dimensional case, we set y = (u, v) ∈ Ω and introduce the labeling yk+(l−1)K = (uk, vl) for k, l = 1, . . . K. i ∼ exp(−(uk − uk′)2) exp(−(vl − vl′)2) for i = k + (l − 1)K and a = k′ + (l′ − 1)K. Hence, we can write I = Tfσ ⊗ Tfσ, where ⊗ is the Kronecker product; the Kronecker product is responsible for the K ×K sub-structure we can observe in figure 3 with K = 10. Note that if we were interested in a n-dimensional retinotopy, then the input matrix could be written as a Kronecker product between n Toeplitz matrices. As previously mentioned, the final connectivity matrix roughly corresponds to the correlation matrix of the input matrix. It turns out that the correlation matrix of I is also a Kronecker product of two Toeplitz matrix generated by a single Gaussian (with a different standard deviation). Thus, the connectivity matrix has the same basic form as the input matrix when za = yi for i = a. The inputs and stable equilibrium points of the simulated system are shown in figure 3. The positions xi of the neurons after multidimensional scaling are shown in figure 4. 5.3.2 Toroıdal retinotopy We now assume that the inputs are uniformly distributed over a two-dimensional torus, i.e. Ω = T2. That is, the input labels za are randomly distributed on the torus. The neuron labels yi are regularly and uniformly distributed on the torus. The inputs and final stable weight matrix of the simulated system are shown in figure 5. The positions xi of the neurons after multidimensional scaling for k = 3 are shown in figure 6, and appear to form a cloud of points distributed on a torus. In order to confirm this, we have used a numerical method from computational co- homology [Zomorodian and Carlsson, 2005] to construct the so–called persistent 17 Figure 4: Positions xi of the neurons after having applied classical multidimen- sional scaling to the final connectivity matrix shown in figure 3 for k = 2 (left) and k = 3 (right). The regular spacing of the neurons for k = 2 shows that the planar structure of the inputs has been recovered, although the corner of the square appear less regular due to boundary effects. In the case k = 3, there is an embedding of the plane into three dimensions; the saddle–like shape accounts for the corner irregularity observed when k = 2. Figure 5: Plot of retinotopic inputs on Ω = T2 (left) and the final connectivity matrix (right) for the system Σ′. The parameters used for this simulation are s(x) = 1+e−4(x−1) , l = 1, µ = 10, ǫ = 0.001, N = 1000, M = 10, 000, σ = 2. 1 18 Figure 6: Left: Positions xi of the neurons for k = 3 after having applied multi- dimensional scaling methods presented in part 5.2 to the final connectivity matrix shown in figure 5. Right: Persistent cohomology barcodes of the cloud of points xi computed using the Jplex software package of [Sexton and Vejdemo-Johansson, ]. (See section 8.4 for a short introduction to persistent cohomology). The triplet of betti numbers (1,2,1) appear stable confirming that the points lie on a 2 dimen- sional torus. 19 cohomology barcodes of the neurons' positions. These determine certain topolog- ical invariants of the underlying space. (See section 8.4 for a short introduction to persistent cohomology and barcodes). The results are also shown in figure 6, and establish that the network has learnt the underlying toroidal geometry of the inputs. 5.4 Links with neuroanatomy The brain is subject to energy constraints which are completely neglected in the above formulation. These constraints most likely have a significant impact on the positions of real neurons in the brain. Indeed, it seems reasonable to assume that the positions and connections of neurons reflect a trade-off between the energy costs of biological tissue and their need to process information effectively. For instance, it has been suggested that a principle of wire length minimization may occur in the brain [Swindale, 1996, Chklovskii et al., 2002]. In our neural mass framework, one may consider that the stronger two neural masses are connected, the larger the number of real axons linking the neurons together. Therefore, mini- mizing axonal length can be read as: the stronger the connection the closer, which is consistent with the convolutional part of the weight matrix. However, the un- derlying geometry of natural inputs is likely to be very high-dimensional, whereas the brain lies in a three-dimensional world. In fact, the cortex is so flat that it is effectively two-dimensional. Hence, the positions of real neurons are different from the positions xi ∈ Rk in a high dimensional vector space; since the cor- tex is roughly two-dimensional, the positions could only be realized physically if k = 2. Therefore, the three-dimensional toric geometry or any higher dimensional structure could not be perfectly implemented in the cortex without the help of non- convolutional long-range connections. Indeed, we suggest that the cortical con- nectivity is made of two parts: i) a local convolutional connectivity corresponding to the convolutional term Gσ in (21), which is consistent with the requirements of energy efficiency, and ii) a non-convolutional connectivity corresponding to the factor M in equation (21), which is required in order to represent various stimulus features. If the cortex were higher-dimensional (k ≫ 2) then M ≡ 1. We illustrate the above claim by considering two examples based on the func- tional anatomy of the primary visual cortex: the emergence of ocular dominance columns and orientation columns, respectively. We proceed by returning to the case of planar retinotopy (section 5.3.1) but now with additional input structure. In the first case, the inputs are taken to be binocular and isotropic, whereas in the second case they are taken to be monocular and anisotropic. The details are 20 presented below. Given a set of prescribed inputs, the network evolves accord- ing to equation (15) and the lateral connections converge to a stable equilibrium. The resulting weight matrix is then projected onto the set of distance matrices for k = 2 (as described in section 5.2) using the stress majorization or SMACOF algorithm for the stress1 cost function as described in [Borg and Groenen, 2005]. We thus assign a position xi ∈ R2 to the ith neuron, i = 1, . . . , N. (Note that the position xi extracted from the weights using multidimensional scaling is distinct from the "physical" position yi of the neuron in the retinocortical plane; the latter determines the center of its receptive field). The convolutional connectivity (Gσ in equation 21) is therefore completely defined: on the planar map of points xi, neurons are isotropically connected to their neighbors; the closer the neurons are the stronger is their convolutional connection. Moreover, since the stimulus fea- ture preferences (orientation, ocular dominance) of each neuron i, i = 1, . . . , N, are prescribed, we can superimpose these feature preferences on to the planar map of points xi. In both examples, we find that neurons with the same ocular or orientation selectivity tend to cluster together (see figures 7 and 8): interpolat- ing these clusters then generates corresponding feature columns. It is important to emphasize that the retinocortical positions yi do not have any columnar struc- ture, that is, they do not form clusters with similar feature preferences. Thus, in contrast to standard developmental models of vertical connections, the columnar structure emerges from the recurrent weights following Hebbian learning and an application of multidimensional scaling. It follows that neurons coding for the same feature tend to be strongly connected; indeed, the multidimensional scaling algorithm has the property that it positions strongly connected neurons close to- gether . Equation (21) also suggests that the connectivity has a non-convolutional part, M, which is a consequence of the low-dimensionality (k = 2). In order to illustrate the structure of the non-convolutional connectivity, we select a neuron i in the plane and draw a link from it at position xi to the neurons at position xj for which M(xi, xj) is maximal. We find that M tends to be patchy, i.e. it connects neurons having the same feature preferences. In the case of orientation, M also tends to be co-aligned, i.e. connecting neurons with similar orientation preference along a vector in the plane of the same orientation. 5.4.1 Ocular dominance columns and patchy connectivity In order to construct binocular inputs, we partition the N neurons into two sets i ∈ {1, . . . , N/2} and i ∈ {N/2 + 1, . . . , N} that code for the left and right eyes, respectively. The ith neuron is then given a retinocortical position yi ∈ [0, 1] × 21 [0, 1], with the yi uniformly distributed across the plane. We do not assume a priori that there exist any ocular dominance columns, that is, neurons with similar retinocortical positions yi do not form clusters of cells coding for the same eye. We then take the ath input to the network to be of the form i = (1 + γ(a))e− (yi−za)2 I (a) , i = (1 − γ(a))e− (yi−za+s)2 I (a) σ′2 σ′2 i = 1, . . . , N/2 , i = N/2 + 1, . . . , N, where s ∈ R2 represents some form of binocular disparity, za and γ(a) are randomly generated from [0, 1]2 and [−1, 1], respectively, see [Bressloff, 2005]. Thus, if γ(a) > 0 (γ(a) < 0) then the corresponding input is predominantly from the left (right) eye. In our simulations we take σ = σ′ = 0.1 and s = 0.005. The results of our simulations are shown in figure 7. In particular, we plot the points xi obtained by performing multidimensional scaling on the final connectivity matrix for k = 2, and superimposing upon this the ocular dominance map obtained by interpolating between clusters of neurons with the same eye preference. We also illustrate the non-convolutional connectivity by linking one selected neuron to the five neurons labeled j it is most strongly connected to (with M(xi, xj) > 1), with i the label of the central neuron. This clearly shows that long–range connections tend to link cells with the same ocular dominance. 5.4.2 Orientation columns and colinear connectivity In order to construct oriented inputs, we partition the N neurons into four groups Σθ corresponding to different orientation preferences θ = {0, π 4 }. Thus, if neuron i ∈ Σθ then its orientation preference is θi = θ. For each group, the neurons are randomly assigned a retinocortical position yi ∈ [0, 1] × [0, 1]. Again, we do not assume a priori that there exist any orientation columns, that is, neurons with similar retinocortical positions yi do not form clusters of cells coding for the same orientation preference. Each cortical input I (a) is generated by convolving a thalamic input consisting of an oriented Gaussian with a Gabor–like receptive field [Miikkulainen et al., 2005]. Let Rθ denote a 2-dimensional rigid body rotation in the plane with θ ∈ [0, 2π). Then 2 , 3π 4 , π i where I (a) i =Z Gi(ξ − yi)Ia(ξ − za)dξ, Gi(ξ) = G0(Rθiξ) 22 (24) (25) Figure 7: Plot of the positions xi of neurons for k = 2 in red (right eye) and green (left eye). We have used an interpolation method to highlight the areas dominated by the right eye in black. These ocular dominance columns have fractal borders which are less regular than those observed in optical imaging experiments. The convolutional connectivity (Gσ in equation (21)) is implicitly described by the po- sition of the neurons: the closer the neurons, the stronger their connections. The strongest components of the non-convolutional connectivity (M in equation (21)) from a central red neuron are also shown by drawing links from this neuron to the target neurons. The color of the link refers to the color of the target neuron. Therefore, we see that it is mainly connected to neurons of its same ocular dom- inance resulting in a patchy distribution. The parameters used for this simulation are s(x) = 1+e−4(x−1) , l = 1, µ = 10, ǫ = 0.01, N = 800 M = 3200. 1 23 and G0(ξ) is the Gabor–like function G0(ξ) = A+e−ξT .Λ−1.ξ − A−e−(ξ−e0)T .Λ−1.(ξ−e0) − A−e−(ξ+e0)T .Λ−1.(ξ+e0) with e0 = (0, 1) and Λ =(cid:18)σlarge 0 0 σsmall(cid:19) . The amplitudes A+, A− are chosen so thatR G0(ξ)dξ = 0. Similarly, the thalamic aξ) with I(ξ) the anisotropic Gaussian input Ia(ξ) = I(Rθ′ I(ξ) = e−ξT .Λ′−1.ξ, Λ′ =(cid:18)σ′ large 0 0 small(cid:19) . σ′ a and za are randomly generated from [0, π) and [0, 1]2 The input parameters θ′ large = 0.266... and respectively. In our simulations we take σlarge = 0.133..., σ′ small = 0.0333.... The results of our simulations are shown in figure 8. σsmall = σ′ In particular, we plot the points xi obtained by performing multidimensional scal- ing on the final connectivity matrix for k = 2, and superimposing upon this the orientation preference map obtained by interpolating between clusters of neurons with the same orientation preference. To avoid border problems we have zoomed on the center on the map. We also illustrate the non-convolutional connectiv- ity by linking one selected neuron to all other neurons for which M is maximal. The patchy, anisotropic nature of the long–range connections is clearly seen. The anisotropic nature of the connections is further quantified in the histogram of fig- ure 9. 6 Discussion In this paper, we have shown how a neural network can learn the underlying geom- etry of a set of inputs. We have considered a fully recurrent neural network whose dynamics is described by a simple non-linear rate equation, together with unsu- pervised Hebbian learning with decay that occurs on a much slower time scale. Although several inputs are periodically presented to the network, so that the re- sulting dynamical system is non-autonomous, we have shown that such a system has a fairly simple dynamics: the network connectivity matrix always converges to an equilibrium point. We have then demonstrated how this connectivity matrix can be expressed as a distance matrix in Rk for sufficiently large k, which can be 24 Figure 8: Plot of the positions xi of neurons for k = 2 obtained by multidi- mensional scaling of the weight matrix. Neurons are clustered in orientation columns represented by the colored areas, which are computed by interpolation. The strongest components of the non-convolutional connectivity (M in equation (21)) from a particular neuron in a yellow area are illustrated by drawing black links from this neuron to the target neurons. Since the yellow color corresponds to an orientation of 3π 4 , the non-convolutional connectivity shows the existence of a co-linear connectivity as exposed in [Bosking et al., 1997]. The parameters used for this simulation are s(x) = 1+e−4(x−1) , l = 1, µ = 10, ǫ = 0.01, N = 900 M = 9000. 1 25 140 120 100 80 60 40 20 0 Figure 9: Histogram of the 5 largest components of the non-convolutional connec- tivity for 80 neurons randomly chosen among those shown in Fig. 8. The abcissa corresponds to the difference in radian between the direction preference of the neuron and the direction of the links between the neuron and the target neurons. This histogram is weighted by the strength of the non-convolutional connectiv- ity. It shows a preference for co-aligned neurons but also a slight preference for perpendicularly-aligned neurons (e.g. neurons of the same orientation but parallel to each other). 26 related to the underlying geometrical structure of the inputs. If the connectivity matrix is embedded in a lower two-dimensional space (k = 2), then the emerg- ing patterns are similar to experimentally observed cortical feature maps. That is, neurons with the same feature preferences tend to cluster together forming cortical columns within the embedding space. Moreover, the recurrent weights decompose into a local isotropic convolutional part, which is consistent with the requirements of energy efficiency, and a longer–range non-convolutional part that is patchy. This suggest a new interpretation of the cortical maps: they correspond to two-dimensional embeddings of the underlying geometry of the inputs. One of the limitations of applying simple Hebbian learning to recurrent corti- cal connections is that it only takes into account excitatory connections, whereas 20% of cortical neurons are inhibitory. Indeed, in most developmental models of feedforward connections, it is assumed that the local and convolutional connec- tions in cortex have a Mexican hat shape with negative (inhibitory) lobes for neu- rons that are sufficiently far from each other. From a computational perspective, it is possible to obtain such a weight distribution by replacing Hebbian learning with some form of covariance learning ([Sejnowski and Tesauro, 1989]). However, it is difficult to prove convergence to a fixed point in the case of the covariance learn- ing rule, and multidimensional scaling method cannot be applied directly unless the Mexican hat function is truncated so that it is invertible. Another limitation of rate-based Hebbian learning is that it does not take into account causality, in contrast to more biologically detailed mechanisms such as spike timing dependent plasticity. The approach taken here is very different from standard treatments of cortical development [Miller et al., 1989, Swindale, 1996], in which the recurrent connec- tions are assumed to be fixed and of convolutional Mexican hat form whilst the feedforward vertical connections undergo some form of correlation-based Heb- bian learning. In the latter case, cortical feature maps form in the physical space of retinocortoical coordinates yi, rather than in the more abstract planar space of points xi obtained by applying multidimensional scaling to recurrent weights undergoing Hebbian learning in the presence of fixed vertical connections. A par- ticular feature of cortical maps formed by modifiable feedforward connections is that the mean size of a column is determined by a Turing-like pattern forming in- stability, and depends on the length scales of the Mexican hat weight function and the two-point input correlations [Miller et al., 1989, Swindale, 1996]. No such Turing mechanism exists in our approach so that the resulting cortical maps tend to be more fractal-like (many length scales) compared to real cortical maps. Nev- ertheless, we have established that the geometrical structure of cortical feature 27 maps can also be encoded by modifiable recurrent connections. This should have interesting consequences for models that consider the joint development of feed- forward and recurrent cortical connections. One possibility is that the embedding space of points xi arising from multidimensional scaling of the weights becomes identified with the physical space of retinocortical positions yi. The emergence of local convolutional structures together with sparser long-range connections would then be consistent with energy efficiency constraints in physical space. Our paper also draws a direct link between the recurrent connectivity of the network and the positions of neurons in some vector space such as R2. In other words, learning corresponds to moving neurons or nodes so that their final position will match the inputs' geometrical structure. Similarly, the Kohonen algorithm [Kohonen, 1990] describes a way to move nodes according to the inputs presented to the network. It also converges toward the underlying geometry of the set of inputs. Although not formally equivalent, it seems that both of these approaches have the same qualitative behaviour. However, our method is more general in the sense that no neighborhood structure is assumed a priori; such a structure emerges via the embedding into Rk. Finally, note that we have used a discrete formalism based on a finite num- ber of neurons. However, the resulting convolutional structure obtained by ex- pressing the weight matrix as a distance matrix in Rk, equation (21), allows us to take an appropriate continuum limit. This then generates a continuous neural field model in the form of an integro-differential equation whose integral kernel is given by the underlying weight distribution. Neural fields have been used increas- ingly to study large–scale cortical dynamics (see [Coombes, 2005] for a review). Our geometrical learning theory provides a developmental mechanism for the for- mation of these neural fields. One of the useful features of neural fields from a mathematical perspective is that many of the methods of partial differential equa- tions can be carried over. Indeed, for a general class of connectivity functions defined over continuous neural fields, a reaction-diffusion equation can be de- rived whose solution approximates the firing rate of the associated neural field [Degond and Mas-Gallic, 1989, Cottet, 1995, Edwards, 1996]. It appears that the necessary connectivity functions are precisely those that can be written in the form (21). This suggests that a network that has been trained on a set inputs with an appropriate geometrical structure behaves as a diffusion equation in a high- dimensional space together with a reaction term corresponding to the inputs. 28 7 Acknowldegments MNG and ODF were partially funded by the ERC advanced grant NerVi nb 227747. MG was partially funded by the r´egion PACA, France. This publica- tion was based on work supported in part by the National Science Foundation (DMS-0813677) and by Award No KUK-C1-013-4 made by King Abdullah Uni- versity of Science and Technology (KAUST). PCB was also partially supported by the Royal Society Wolfson Foundation. 8 Appendix 8.1 Proof of the convergence to the symmetric attractor A We need to prove the 2 points: (i) A is an invariant set, and (ii) for all Y(0) ∈ RN ×M × RN ×N , Y(t) converges to A as t → +∞. Since RN ×N is the direct sum of the set of symmetric connectivities and the set of anti-symmetric connectivies, we write W (t) = WS(t) + WA(t), ∀t ∈ R+, where WS is symetric and WA is anti-symetric. (i) In (15), the right hand side of the equation for W is symmetric. Therefore, if ∃t1 ∈ R+ such that WA(t1) = 0, then W remains in A for t ≥ t1. (ii) Projecting the expression for W in equation (15) on to the anti-symmetric component leads to dWA dt = −ǫµWA(t) (26) whose solution is WA(t) = WA(0) exp(−ǫµt), ∀t ∈ R+. Therefore, 0. The system converges exponentially to A. lim t→+∞ WA(t) = 8.2 Proof of theorem 4.1 Consider the following Lyapunov function (see equation (16)) where µ = µM, such that if W = WS + WA, where WS is symmetric and WA is anti-symmetric. E(U, W ) = − 1 2 hU, W · Ui − hI, Ui + h1, S−1(cid:0)U(cid:1)i + − ∇E(U, W ) =(cid:18)WS · U + I − S−1(cid:0)U(cid:1) (cid:19) U · U T − µW 29 µ 2 kW k2, (27) (28) Therefore, writing the system Σ′, equation (15), as dY dt = γ(cid:18)WS · S(V) + I − S−1(cid:0)S(V)(cid:1) S(V) · S(V)T − µW (cid:19) + γ(cid:18)WA.S(V) 0 (cid:19) , where Y = (V, W )T , we see that dY dt = −γ(cid:18)∇E(cid:0)σ(V, W )(cid:1)(cid:19) + Γ(t) (29) where γ(V, W )T = (V, ǫW/M)T , σ(V, W ) = (S(V), W ) and Γ : R+ → H such that kΓk → 0 exponentially (because the system converges to A). It follows that the time derivative of E = E ◦ σ along trajectories is given by: t→+∞ d E dt =(cid:28)∇ E, dY dt(cid:29) =(cid:28)∇V E, dV dt(cid:29) +(cid:28)∇W E, dW dt (cid:29). (30) Substituting equation (29) then yields d E dt = −(cid:28)∇ E, γ(cid:0)∇E ◦ σ(cid:1)(cid:29) +(cid:28)∇ E, Γ(t)(cid:29) } {z M(cid:28)∇W E ◦ σ, ∇W E ◦ σ(cid:29) + Γ(t). = −(cid:28)S′(V)∇U E ◦ σ, ∇U E ◦ σ(cid:29) − Γ(t) ǫ (31) We have used the chain–rule of differentiation, whereby ∇V ( E) = ∇V (E ◦ σ) = S′(V)∇U E ◦ σ, and S′(V)∇U E (without dots) denotes the Hadamard (term by term) product, that is, [S′(V)∇U E]ia = s′(V (a) i ) ∂E ∂U (a) i Note that Γ → t→+∞ 0 exponentially because ∇ E is bounded, and S′(V) > 0 because the trajectories are bounded. Thus, there exists t1 ∈ R+ such that ∀t > t1, ∃k ∈ R∗ + such that d E dt ≤ −kk∇E ◦ σk2 ≤ 0. (32) As in [Cohen and Grossberg, 1983] and [Dong and Hopfield, 1992], we apply the Krasovskii-LaSalle invariance principle [Khalil and Grizzle, 1996]. We check that: 30 • E is lower bounded. Indeed, V and W are bounded. Given that I and S are also bounded it is clear that E is bounded. d E dt • is negative semidefinite on the trajectories as shown in equation (32). Then the invariance principle tells us that the solutions of the system Σ′ approach the set M = nY ∈ H : H : ∇E ◦ σ = 0o. Since d E dt dY dt (Y) = 0o. Equation (32) implies that M = nY ∈ = −γ(cid:16)∇E ◦ σ(cid:17) and γ 6= 0 everywhere, M consists of the equilibrium points of the system. This completes the proof. 8.3 Proof of theorem 4.2 Denote the right–hand side of system Σ′, equation (15) by F (V, W ) =( ǫ −V + W · S(cid:0)V(cid:1) + I M(cid:0)S(V).S(V)T − µMW(cid:1) ǫ The fixed points satisfy the condition F (V, W ) = 0 which immediately leads to equations (19). Let us now check the linear stability of this system. The differen- tial of F at V ∗, W ∗ is where S′(V ∗)Z denotes a Hadamard product, that is, [S′(V ∗)Z]ia = s′(V ∗ i −Z + W ∗ ·(cid:0)S′(V ∗)Z(cid:1) + J · S(V ∗) M(cid:16)(cid:0)S′(V ∗)Z(cid:1) · S(V ∗)T + S(V ∗) ·(cid:0)S′(V ∗)Z(cid:1)T − µMJ(cid:17),! dF(V ∗,W ∗)(Z, J) = J(cid:19) = Assume that there exist λ ∈ C∗, (Z, J) ∈ H such that dF(V ∗,W ∗)(cid:18)Z J(cid:19). Taking the second component of this equation and computing the dot λ(cid:18)Z (a))Z (a) . i product with S(V ∗) leads to (λ + ǫµ)J · S = ǫ M (cid:0)(S′Z) · ST · S + S · (S′Z)T · S(cid:1) where S = S(V ∗), S′ = S′(V ∗). Substituting this expression in the first equation leads to M(λ + ǫµ)(λ + 1)Z = ( λ µ + ǫ)S · ST · (S′Z) + ǫ(S′Z) · ST · S + ǫS · (S′Z)T · S (33) 31 Observe that setting ǫ = 0 in the previous equation leads to an eigenvalue equation for the membrane potential only: (λ + 1)Z = 1 µM S · ST · (S′Z). µM(cid:0)S · ST(cid:1), this equation implies that λ + 1 is an eigenvalue of Since W ∗ = 1 the operator X 7→ W ∗.(S′X). The magnitudes of the eigenvalues are always smaller than the norm of the operator. Therefore, we can say that if 1 > kW ∗kS′ m then all the possible eigenvalues λ must have a negative real part. This sufficient condition for stability is the same as in [Faugeras et al., 2008]. It says that fixed points sufficiently close to the origin are always stable. Let us now consider the case ǫ 6= 0. Recall that Z is a matrix. We now "flatten" Z by storing its rows in a vector called Zrow. We use the following result in [Brewer, 1978]: the matrix notation of operator X 7→ A·X ·B is A⊗BT , where ⊗ is the Kronecker product. In this formalism the previous equation becomes M(λ+ǫµ)(λ+l)Zrow =(cid:18)( λ µ +ǫ)S·ST ⊗Id+ǫId⊗ST ·S+ǫS⊗ST(cid:19)·(S′Z)row (34) where we assume that the Kronecker product has the priority over the dot product. We focus on the linear operator O defined by the right hand side and bound its kXk which is equal norm. Note that we use the following norm kW k∞ = supX to the largest magnitude of the eigenvalues of W . kW.Xk kOk∞ ≤ S′ m(cid:18) λ µ kS · ST ⊗ Idk∞ + ǫkS · ST ⊗ Idk∞ + ǫkId ⊗ ST · Sk∞ + ǫkS ⊗ ST k∞(cid:19). (35) Define, νm to be the magnitude of the largest eigenvalue of W ∗ = 1 µM (S · ST ). First, note that S · ST and ST · S have the same eigenvalues (µM)νi but different eigenvectors denoted by ui for S · ST and vi for ST · S. In the basis set spanned by the ui ⊗ vj, we find that S · ST ⊗ Id and Id ⊗ ST · S are diagonal with (µM)νi as eigenvalues. Therefore, kS·ST ⊗Idk∞ = (µM)νm and kId⊗ST ·Sk∞ = (µM)νm. Moreover, observe that (ST ⊗S)T ·(ST ⊗S)·(ui⊗vj) = (S·ST ·ui)⊗(ST ·S·vj) = (µM)2νiνj ui⊗vj (36) 32 Therefore, (ST ⊗ S)T · (ST ⊗ S) = (µM)2diag(νiνj). In other words, ST ⊗ S is the composition of an orthogonal operator (i.e. an isometry) and a diagonal matrix. Immediately, it follows that kST ⊗ Sk ≤ (µM)νm. Compute the norm of equation (34) (λ + ǫµ)(λ + 1) ≤ S′ m(λ + 3ǫµ)νm. (37) Define fǫ : C → R such that fǫ(λ) = (λ + ǫµ)(λ + 1) − (λ + 3ǫµ)S′ mνm. We want to find a condition such that fǫ(C+) > 0, where C+ is the right half complex plane. This condition on ǫ, µ, νm, and S′ m will be a sufficient condition for linear stability. Indeed, under this condition we can show that only eigenvalues with a negative real part can meet the necessary condition (37). Complex number of the right half plane cannot be eigenvalues and thus the system is stable. The case ǫ = 0 tells us that f0(C+) > 0 if 1 > S′ mνm, compute ∂fǫ ∂ǫ (λ) = µ(ℜ(λ) + µǫ) (λ + 1) (λ + ǫµ) − 3µS′ mνm If 1 ≥ ǫµ, which is most probably true given that ǫ << 1, then (λ+1) Assuming λ ∈ C+ leads to: (λ+ǫµ) ≥ 1. ∂fǫ ∂ǫ (λ) ≥ µ(µǫ − 3S′ mνm) ≥ µ(1 − 3S′ mνm) Therefore, the condition 3S′ fǫ(C+) > 0. mνm < 1, which implies S′ mνm < 1, and leads to 8.4 A very short introduction to computational cohomology In algebraic topology, topological spaces (which are continuous objects) can be classified by roughly counting their number of holes. This coordinate-invariant description of a topological space is called its homology (or cohomology, the difference between them is beyond the scope of this paper). In fact the homology can be summarized by giving the betti numbers of the topological state. The sequence of betti number is made of positive integers. The first three betti numbers have the following definition: the first is the number of connected components, the second is the number of two-dimensional or "circular" holes and the third is the number of 3-dimensional holes or "voids". See [Hatcher, 2002] for a more rigorous approach. 33 However, in the example of toroidal retinotopy (see section 5.3.2 and figure 6). we are dealing with a discrete cloud of points. Therefore, one needs to extend the definition of the betti numbers to discrete objects in order to find the underly- ing topology of the space within which the points are distributed. This is called computational or persistent cohomology. One reconstructs the topological space by considering balls of a given radius centered on each point in the cloud. For each radius (the abscissa of the right picture of figure 6), one can compute the betti numbers of the resulting topological space. A barcode graph, e.g. the right picture of figure 6, is constructed by drawing a horizontal bar for each connected component, 2-dimensional hole or 3-dimenional void etc. This is done for a range of radii. Finding the persistent cohomology of a cloud of points consists in ob- serving the set of betti numbers that are stable through a significantly large range of radii. One then assumes that the points most likely lie on a topological space of a given homology if the corresponding betti numbers are stable enough. See [Zomorodian and Carlsson, 2005] for details. In section 5.3.2, we used the Jplex software package of [Sexton and Vejdemo-Johansson, ] to compute the barcodes of the points corresponding to learning from inputs uni- formly distributed over a 2-dimensional torus. We used 200 landmarks spread according to the maxminlandmark method to build simplices in Jplex, which re- turned the maximum radius (beyond which all the betti numbers except the first vanish) we used in the simulation. In figure 6, we see that for a wide range of radii the triplet (1, 2, 1) is stable. This corresponds to a 2 dimensional torus. References [Amari, 1998] Amari, S. (1998). Natural gradient works efficiently in learning. Neural computation, 10(2):251–276. [Amari et al., 1992] Amari, S., Kurata, K., and Nagaoka, H. (1992). Information IEEE Transactions on Neural Networks, geometry of Boltzmann machines. 3(2):260–271. [Bartsch and Van Hemmen, 2001] Bartsch, A. and Van Hemmen, J. (2001). Combined Hebbian development of geniculocortical and lateral connectivity in a model of primary visual cortex. Biological Cybernetics, 84(1):41–55. 34 [Bi and Poo, 2001] Bi, G. and Poo, M. (2001). Synaptic modification by cor- related activity: Hebb's postulate revisited. Annual review of neuroscience, 24:139. [Bienenstock et al., 1982] Bienenstock, E., Cooper, L., and Munro, P. (1982). Theory for the development of neuron selectivity: orientation specificity and binocular interaction in visual cortex. J Neurosci, 2:32–48. [Borg and Groenen, 2005] Borg, I. and Groenen, P. (2005). Modern multidimen- sional scaling: Theory and applications. Springer Verlag. [Bosking et al., 1997] Bosking, W., Zhang, Y., Schofield, B., and Fitzpatrick, D. (1997). Orientation selectivity and the arrangement of horizontal connections in tree shrew striate cortex. Journal of neuroscience, 17(6):2112. [Bressloff, 2005] Bressloff, P. (2005). Spontaneous symmetry breaking in self– organizing neural fields. Biological Cybernetics, 93(4):256–274. [Bressloff et al., 2001] Bressloff, P., Cowan, J., Golubitsky, M., Thomas, P., and Wiener, M. (2001). Geometric visual hallucinations, euclidean symmetry and the functional architecture of striate cortex. Phil. Trans. R. Soc. Lond. B, 306(1407):299–330. [Bressloff and Cowan, 2003] Bressloff, P. C. and Cowan, J. D. (2003). A spher- ical model for orientation and spatial frequency tuning in a cortical hypercol- umn. Philosophical Transactions of the Royal Society B. [Brewer, 1978] Brewer, J. (1978). Kronecker products and matrix calculus in system theory. IEEE Transactions on Circuits and Systems, 25(9). [Chklovskii et al., 2002] Chklovskii, D., Schikorski, T., and Stevens, C. (2002). Wiring optimization in cortical circuits. Neuron, 34(3):341–347. [Chossat and Faugeras, 2009] Chossat, P. and Faugeras, O. (2009). Hyperbolic planforms in relation to visual edges and textures perception. PLoS Computa- tional Biology, 5(12):367–375. [Cohen and Grossberg, 1983] Cohen, M. and Grossberg, S. (1983). Absolute sta- bility of global pattern formation and parallel memory storage by competitive neural networks. In IEEE Transactions on Systems, Man, and Cybernetics, SMC-13, pages 815–826. 35 [Coifman et al., 2005] Coifman, R., Maggioni, M., Zucker, S., and Kevrekidis, I. (2005). Geometric diffusions for the analysis of data from sensor networks. Current opinion in neurobiology, 15(5):576–584. [Coombes, 2005] Coombes, S. (2005). Waves, bumps, and patterns in neural field theories. Biological Cybernetics, 93(2):91–108. [Cottet, 1995] Cottet, G. (1995). Neural networks: Continuous approach and applications to image processing. Journal of Biological Systems, 3:1131–1139. [Dayan and Abbott, 2001] Dayan, P. and Abbott, L. (2001). Theoretical Neuro- science : Computational and Mathematical Modeling of Neural Systems. MIT Press. [Degond and Mas-Gallic, 1989] Degond, P. and Mas-Gallic, S. (1989). The Weighted Particle Method for Convection-Diffusion Equations. Part 1: The Case of an Isotropic Viscosity. Mathematics of Computation, pages 485–507. [Dong and Hopfield, 1992] Dong, D. and Hopfield, J. (1992). Dynamic prop- erties of neural networks with adapting synapses. Network: Computation in Neural Systems, 3(3):267–283. [Edwards, 1996] Edwards, R. (1996). Approximation of neural network dynam- ics by reaction-diffusion equations. Mathematical methods in the applied sci- ences, 19(8):651–677. [Faugeras et al., 2008] Faugeras, O., Grimbert, F., and Slotine, J.-J. (2008). Abo- lute stability and complete synchronization in a class of neural fields models. SIAM J. Appl. Math, 61(1):205–250. [Foldi´ak, 1991] Foldi´ak, P. (1991). Learning invariance from transformation se- quences. Neural Computation, 3(2):194–200. [Geman, 1979] Geman, S. (1979). Some averaging and stability results for ran- dom differential equations. SIAM J. Appl. Math, 36(1):86–105. [Gerstner and Kistler, 2002] Gerstner, W. and Kistler, W. M. (2002). Mathemati- cal formulations of hebbian learning. Biological Cybernetics, 87:404–415. [Hatcher, 2002] Hatcher, A. (2002). Algebraic topology. Cambridge Univ Pr. 36 [Hebb, 1949] Hebb, D. (1949). The organization of behavior: a neuropsycholog- ical theory. Wiley, NY. [Hubel and Wiesel, 1977] Hubel, D. H. and Wiesel, T. N. (1977). Functional architecture of macaque monkey visual cortex. Proc. Roy. Soc. B, 198:1–59. [Khalil and Grizzle, 1996] Khalil, H. and Grizzle, J. (1996). Nonlinear systems. Prentice hall Upper Saddle River, NJ. [Kohonen, 1990] Kohonen, T. (1990). The Self-Organizing Map. Proceedings of the IEEE, 78(9). [Lawlor and Zucker, 2010] Lawlor, M. and Zucker, S. (2010). Third-Order Edge Statistics Reveal Curvature Dependency. In Snowbird workshop on learning. [Miikkulainen et al., 2005] Miikkulainen, R., Bednar, J., Choe, Y., and Sirosh, J. (2005). Computational Maps in the Visual Cortex. Springer, New York. [Miller, 1996] Miller, K. (1996). Synaptic economics: competition and coopera- tion in synaptic plasticity. Neuron, 17:371–374. [Miller and MacKay, 1996] Miller, K. and MacKay, D. (1996). The role of con- straints in hebbian learning. Neural Comp, 6:100–126. [Miller et al., 1989] Miller, K. D., Keller, J. B., and Stryker, M. P. (1989). Ocular dominance column development: analysis and simulation. Science, 245:605– 615. [Oja, 1982] Oja, E. (1982). A simplified neuron model as a principal component analyzer. J. Math. Biology, 15:267–273. [Ooyen, 2001] Ooyen, A. (2001). Competition in the development of nerve con- nections: a review of models. Network: Computation in Neural Systems, 12(1):1–47. [Petitot, 2003] Petitot, J. (2003). The neurogeometry of pinwheels as a sub- Riemannian contact structure. Journal of Physiology-Paris, 97(2-3):265–309. [Sejnowski and Tesauro, 1989] Sejnowski, T. and Tesauro, G. (1989). The Hebb rule for synaptic plasticity: algorithms and implementations. Neural models of plasticity: Experimental and theoretical approaches, pages 94–103. 37 [Sexton and Vejdemo-Johansson, ] Sexton, and Johansson, complex http://comptop.stanford.edu/programs/jplex/. M. H. JPlex simplicial Vejdemo- library. [Swindale, 1996] Swindale, N. (1996). The development of topography in the visual cortex: a review of models. Network: Computation in Neural Systems, 7(2):161–247. [Takeuchi and Amari, 1979] Takeuchi, A. and Amari, S. (1979). Formation of topographic maps and columnar microstructures in nerve fields. Biological Cybernetics, 35(2):63–72. [Tikhonov, 1952] Tikhonov, A. (1952). Systems of differential equations with small parameters multiplying the derivatives. Matem. sb, 31(3):575–586. [Verhulst, 2007] Verhulst, F. (2007). Singular perturbation methods for slow–fast dynamics. Nonlinear Dynamics, 50(4):747–753. [Wallis and Baddeley, 1997] Wallis, G. and Baddeley, R. (1997). Optimal, un- supervised learning in invariant object recognition. Neural computation, 9(4):883–894. [Zomorodian and Carlsson, 2005] Zomorodian, A. and Carlsson, G. (2005). Computing persistent homology. Discrete and Computational Geometry, 33(2):249–274. 38
1003.5557
1
1003
2010-03-29T14:51:39
A ratchet mechanism for amplification in low-frequency mammalian hearing
[ "q-bio.NC", "physics.bio-ph" ]
The sensitivity and frequency selectivity of hearing result from tuned amplification by an active process in the mechanoreceptive hair cells. In most vertebrates the active process stems from the active motility of hair bundles. The mammalian cochlea exhibits an additional form of mechanical activity termed electromotility: its outer hair cells (OHCs) change length upon electrical stimulation. The relative contributions of these two mechanisms to the active process in the mammalian inner ear is the subject of intense current debate. Here we show that active hair-bundle motility and electromotility can together implement an efficient mechanism for amplification that functions like a ratchet: sound-evoked forces acting on the basilar membrane are transmitted to the hair bundles whereas electromotility decouples active hair-bundle forces from the basilar membrane. This unidirectional coupling can extend the hearing range well below the resonant frequency of the basilar membrane. It thereby provides a concept for low-frequency hearing that accounts for a variety of unexplained experimental observations from the cochlear apex, including the shape and phase behavior of apical tuning curves, their lack of significant nonlinearities, and the shape changes of threshold tuning curves of auditory nerve fibers along the cochlea. The ratchet mechanism constitutes a general design principle for implementing mechanical amplification in engineering applications.
q-bio.NC
q-bio
A ratchet mechanism for amplification in low-frequency mammalian hearing Tobias Reichenbach and A. J. Hudspeth ∗Howard Hughes Medical Institute and Laboratory of Sensory Neuroscience, The Rockefeller University, New York, New York 10065-6399, U.S.A. ∗ Submitted to Proceedings of the National Academy of Sciences of the United States of America 0 1 0 2 r a M 9 2 ] . C N o i b - q [ 1 v 7 5 5 5 . 3 0 0 1 : v i X r a The sensitivity and frequency selectivity of hearing result from tuned amplification by an active process in the mechanoreceptive hair cells. In most vertebrates the active process stems from the active motil- ity of hair bundles. The mammalian cochlea exhibits an additional form of mechanical activity termed electromotility: its outer hair cells (OHCs) change length upon electrical stimulation. The rela- tive contributions of these two mechanisms to the active process in the mammalian inner ear is the subject of intense current debate. Here we show that active hair-bundle motility and electromotility can together implement an efficient mechanism for amplification that functions like a ratchet: sound-evoked forces acting on the basilar membrane are transmitted to the hair bundles whereas electromotil- ity decouples active hair-bundle forces from the basilar membrane. This unidirectional coupling can extend the hearing range well below the resonant frequency of the basilar membrane. It thereby provides a concept for low-frequency hearing that accounts for a variety of unexplained experimental observations from the cochlear apex, in- cluding the shape and phase behavior of apical tuning curves, their lack of significant nonlinearities, and the shape changes of threshold tuning curves of auditory nerve fibers along the cochlea. The ratchet mechanism constitutes a general design principle for implementing mechanical amplification in engineering applications. auditory system cochlea hair cell T he mammalian cochlea acts as a frequency analyzer in which high frequencies are detected at the organ's base and low frequencies at more apical positions. This frequency mapping is thought to be achieved by a position-dependent resonance of the elastic basilar membrane separating two fluid- filled compartments (Fig. 1A) [1, 2, 3]. When sound evokes a pressure wave that displaces the basilar membrane, the re- sultant traveling wave gradually increases in amplitude as it progresses to the position where the basilar membrane's reso- nant frequency coincides with that of the stimulus. Aided by mechanical energy provided by the active process, the wave peaks at a characteristic place slightly before the resonant po- sition and then declines sharply (Fig. 1B). This mechanism is termed critical-layer absorption [1], for a wave cannot travel beyond its characteristic position on the basilar membrane, but peaks and dissipates most of its energy there. The mech- anism displays scale invariance: different stimulation frequen- cies induce traveling waves that display a common, strongly asymmetric form upon rescaling of the amplitude and spatial coordinate [4, 5]. Two important aspects of the cochlea's mechanics remain problematical. First, the basilar membrane's resonant fre- quency apparently cannot span the entire range of audible fre- quencies. Experimental measurements of basilar-membrane stiffness suggest that high-frequency resonances are feasible but low-frequency ones are inaccessible [6]. This result accords with the analysis of threshold tuning curves for auditory- nerve fibers [7, 8, 9] and measurements of basilar-membrane displacement [8, 10, 11, 12], both of which indicate that a peaked traveling wave occurs for high-frequency stimulation but not for low-frequency stimulation. Indeed, threshold tun- ing curves of auditory-nerve fibers [7, 8, 9] show that high- frequency curves are scale-invariant and possess a sharp cut- off at frequencies above their characteristic frequencies, re- 1 -- 16 flecting the mechanism of critical-layer absorption. However, low-frequency curves lack this sharp cutoff and, in contra- diction of the expectation for critical-layer absorption, are instead characterized by an approximately symmetric shape around their characteristic frequencies. Recent experiments in the chinchilla have shown that the shape change between the tuning curves of high- and low-frequency fibers occurs between two crossover frequencies of about 5 kHz and 1.5 kHz [8, 9]. Measurements of basilar-membrane displacement yield similar conclusions. Experiments from the cochlear base confirm the existence at high frequencies of peaked travel- ing waves that result in strongly asymmetric tuning curves and a pronounced nonlinearity at the characteristic frequen- cies [13, 14, 8, 12]. Apical measurements of basilar-membrane displacement, however, produce symmetric tuning curves that lack a sharp cutoff for frequencies higher than the character- istic frequency [10, 15, 11]. Moreover, only small nonlinearity has been measured [15, 16, 10], raising the question whether amplification occurs at the apex. All available experimental results therefore indicate that low-frequency hearing does not function through critical-layer absorption but must rely on another mechanism to achieve frequency selectivity. The second key uncertainty is the nature of the active process that operates in the mammalian cochlea. In addi- tion to amplifying weak signals, the active process produces increased frequency selectivity, compressive nonlinearity, and spontaneous otoacoustic emission. In the mammalian cochlea amplification is provided by specialized OHCs located in the organ of Corti along the basilar membrane (Fig. 2A). Each OHC displays two forms of motility. Like those in the hear- ing organs of other vertebrates [17, 18], the hair bundle of an OHC can produce mechanical force [17, 18, 19, 20, 21] (see [22] for a review). But an OHC also exhibits electro- motility: when its membrane potential changes as a result of sound-evoked hair-bundle deflection, the entire cell under- goes a length change owing to conformational rearrangement of the membrane protein prestin [23, 24, 25]. Direct measure- ments of the respective roles of the two forms of motility are complicated, for it is difficult to determine the micromechan- ical responses of the organ of Corti while the cochlea remains intact. Here we have taken a theoretical approach to investigate a possible mechanism for amplification by the synergistic in- terplay of active hair-bundle motility and electromotility. We show that they can operate together to achieve low-frequency selectivity in the absence of basilar-membrane resonance and critical-layer absorption. The model advances theoretical un- derstanding of cochlear mechanics in three ways: it reproduces previous theoretical results insofar as they coincide with ex- periments; it accounts naturally for a variety of unexplained findings from the cochlear apex; and it makes robust, experi- mentally testable predictions. Fig. 1. Principles of cochlear mechanics. (A) In a schematic diagram of the mam- malian cochlea, the basilar membrane (BM) is displaced by sound stimuli acting on the stapes (top left). (B) In the classical theory of cochlear mechanics, sound evokes a pressure wave that causes a longitudinal traveling wave of basilar-membrane dis- placement (thick line). The motion of the basilar membrane and the displacements of the associated hair bundles are approximately equal. As the wave approaches the position where its frequency matches the basilar membrane's resonant frequency, the wave's amplitude (thin line) increases and its wavelength and velocity decline. The wave peaks at a characteristic place slightly before the resonant position and then de- clines sharply, yielding a strongly asymmetric envelope of the traveling wave (shading). Experiments confirm this behavior in the basal, high-frequency part of the cochlea. (C) We propose an alternative theory for the cochlea's mechanics at low frequencies. The basilar membrane near the cochlear apex does not resonate, but the traveling wave on the basilar membrane propagates along the entire cochlea without a strong variation in amplitude, wavelength, and velocity (black). However, the interplay of electromotility and active hair-bundle motility fosters an independent resonance of the complex formed by the hair bundles, reticular lamina, and tectorial membrane. The hair-bundle displacement (red) at the characteristic place can therefore exhibit an approximately symmetric peak, exceeding basilar-membrane motion by orders of magnitude. Results The Ratchet Mechanism. Consider a transverse element of the cochlear partition comprising the basilar membrane, an inner hair cell (IHC) and three OHCs, and the overlying tectorial membrane (Fig. 2A). When a sound-evoked force displaces the basilar membrane upward, the resultant shearing motion between the tectorial membrane and the top of the OHCs, or reticular lamina, deflects the hair bundles in the positive direction. Active hair-bundle motility in the OHCs increases the amplitude of deflection. Without electromotility, this ad- ditional displacement would couple back to the basilar mem- brane and augment the movement there. If electromotility is adjusted such that the OHCs elongate just as much as the tectorial membrane and reticular lamina move upward, how- ever, the basilar membrane does not experience the active force and thus undergoes no additional displacement (see also Movie S1). A mathematical formulation of this amplification mech- anism clarifies its operation. Consider a two-mass model in which the two degrees of freedom are the motion of the basilar membrane, XBM, and that of the complex formed by 2 Fig. 2. The ratchet mechanism. (A) The organ of Corti rests upon the basilar membrane (BM). Three OHCs are connected to Deiters' cells (DC), which together couple the basilar membrane to the reticular lamina (dark green, top of the OHCs) and through the hair bundles to the overlying tectorial membrane (TM). Sound-evoked external forces (black arrow) displace the basilar membrane, here upwards, and pro- duce shearing (black arrow) of the hair bundles of OHCs (red asterisk) and the inner hair cell (IHC) (cyan asterisk). Two forms of motility underlie the active process: active hair-bundle motility (single-headed red arrow) and membrane-based electro- motility (double-headed red arrow). (B) The two fundamental degrees of freedom are the basilar-membrane displacement XBM and the displacement XHB of the hair-bundle complex (circle), which comprises the hair bundles, reticular lamina (RL), and tectorial membrane. Coupling stems from the impedance ZD of the combined OHCs and Deiters' cells as well as the impedance ZC of the remaining organ of Corti. (C) In the ratchet mechanism displacements of the basilar membrane caused by external forces are communicated to the hair-bundle complex. Internal forces (red arrow) in the hair bundles increase the shearing motion, which decouples from basilar- membrane displacement through appropriate length changes of the OHCs (dotted red arrow). For an animated representation of the model, see Movie S1. the hair bundles, reticular lamina, and tectorial membrane, XHB (Fig. 2B). A sound stimulus of frequency f produces an oscillating external force Fext(t) = Fexte2πif t + c.c. act- ing on the basilar membrane, in which c.c denotes the com- plex conjugate. The evoked oscillations of the basilar mem- brane as well as the hair-bundle complex occur predomi- nantly at the same frequency, XBM = XBMe2πif t + c.c. and XHB = XHBe2πif t + c.c.. Internal forces Fint arise within hair bundles, where they provide negative damping and intro- duce nonlinearities (Supporting Information). The cell bod- ies of OHCs can be described as piezoelectric elements [26]. For small, physiologically relevant motions their electrically evoked displacement XEE = XEEe2πif t + c.c. is proportional to the hair-bundle displacement, which triggers changes in the membrane potential, so XEE = −α XHB + c.c. with a com- plex mechanomotility coefficient α [27]. The hair-bundle and basilar-membrane displacements then depend linearly on the forces: . [ 1 ] (cid:18) XHB (cid:19) XBM A = (cid:18) Fint (cid:19) Fext 1 2πif The matrix A contains the impedances ZHB and ZBM of the hair-bundle complex and the basilar membrane as well as the coupling impedances ZD and ZC (Fig. 2B): (cid:18)ZHB + (1 + α)ZD + ZC −ZD − ZC (cid:19) A = −(1 + α)ZD − ZC ZBM + ZD + ZC . [ 2 ] high frequencylow frequencyBMApexBaseADisplacementBDisplacementRelative position01CA*DC*TMIHCBMOHCBCHair−bundle.TMcomplexRLBMXBMXHBXEEZDZCcoupledecoupleExternal force.Internal forceHBBM A key feature of this relation is that the matrix ele- ment A21, which describes the coupling of the internal force to the basilar membrane, vanishes at a critical value α∗ ≡ −1 − ZC/ZD. At the same time, the coupling of the external force to the hair bundle, which is represented by element A12, remains nonzero. For the critical value α∗ the displacements become XHB = XBM = 1 2πif ZHB Fint + 1 2πif (ZBM + ZD + ZC) Fext . ZD + ZC 2πif ZHB(ZBM + ZD + ZC) Fext , [ 3 ] The hair-bundle displacement depends on both the internal and external forces, whereas the basilar membrane experiences only the external force. In other words, the sound-evoked ex- ternal force on the basilar membrane is transmitted to the hair bundle, but the internal force from active hair-bundle motility does not feed back onto the motion of the basilar membrane (Fig. 2C). This symmetry-breaking mode has the characteristics of a ratchet in the sense that information flows unidirectionally: information applied as a force against the basilar membrane can be detected in the form of hair-bundle displacement, but force acting on the hair bundle cannot be detected at the basilar membrane. The basilar-membrane dis- placement equals that occuring when the hair-bundle complex is fixed at XHB = 0 and electromotility is absent (α = 0). The ratchet mechanism therefore resembles an ideal operational amplifier that neither feeds back on nor draws energy from the input [28]. The Mechanomotility Coefficient α. The magnitude and phase of the critical value α∗ depend on the coupling impedances ZC and ZD. Under realistic assumptions ZC is similar to or smaller than ZD and α∗ is therefore near unity. Such a value for the mechanomotility coefficient α has been measured experimentally for apical OHCs [27]. The phase of α is controlled by the complex network of ion channels that reg- ulates the membrane potential depending on the hair-bundle displacement [27, 29]. Although experiments on the phase of α are not available, theoretical considerations confirm that OHCs have sufficient flexibility through ion-channel regula- tion to adjust the phase of α to a variety of values, and in particular to the phase required for α∗ (Supporting Informa- tion). A Concept for Low-Frequency Hearing. As its most impor- tant characteristic, the ratchet mechanism can explain fre- quency selectivity near the cochlear apex in the absence of a peaked traveling wave. The resonant frequency of the hair- bundle complex is determined solely by the impedance ZHB and the internal forces, and is therefore defined by the proper- ties of the hair bundles and tectorial membrane (Equation 3) [30, 31, 32]. In particular, hair-bundle resonance can occur in the absence of basilar-membrane resonance. The ratchet mechanism therefore permits the resonant frequency of the hair bundles to follow a logarithmic law along the cochlea, whereas the resonant frequency of the basilar membrane re- mains significantly greater near the apex (Fig. 3B) [6]. Further evidence points to the occurrence of the ratchet mechanism at the cochlear apex but not at the base. First, the basilar membrane at the base is narrow and presumably tuned to the characteristic frequencies at which auditory- nerve fibers are most sensitive (Fig. 3B). Amplification of basilar-membrane motion there is feasible and need not be avoided. Indeed, experiments have demonstrated a strong compressive nonlinearity of basilar-membrane motion at high 3. Cochlear model. Fig. (A) The ratchet mechanism operates when the mechanomotility coefficient α (green) coincides with the critical value α∗ (grey). Although electromotility is negligible to the basal side of fH, it underlies the ratchet mechanism apical to the position of fL. (B) The resonant frequency of the hair- bundle complex (HB, red) agrees with that of the basilar membrane (BM, blue) only for frequencies above fL. (C) A high-frequency sound stimulus (f1 = 8 kHz) in- duces a traveling wave that peaks in the basal region. The displacements of the hair bundles (red) coincide with that of the basilar membrane (blue). Elimination of active hair-bundle motility decreases the sensitivity by a factor of 90, 000 (green, hair bundles; black, basilar membrane) indicative of a strong nonlinearity. A low- frequency stimulus (f2 = 200 Hz) triggers a traveling wave that does not peak on the basilar membrane, but the hair-bundle displacement exhibits a resonance enabled by the ratchet mechanism (same color code). Without active hair-bundle motion the hair-bundle displacement decreases by a factor of only ten indicative of a weak non- linearity. (D) The phase of the basilar-membrane displacement for f1 has a strongly increasing slope near the resonant position and thus shows a wave traveling to the resonant position but not beyond. For f2 the slope of the phase remains almost constant, corresponding to a wave traveling beyond the characteristic place. frequencies. However, the apex exhibits a wider basilar mem- brane that experiences stronger viscous forces and whose resonant frequencies deviate from the characteristic frequen- cies [6]. Amplification of basilar-membrane motion near the apex would therefore be highly inefficient. The absence of a strong compressive nonlinearity in experimental measure- ments of basilar-membrane displacement in the apical region confirms the lack of basilar-membrane amplification there. Next, the membrane time constant of OHCs restricts elec- tromotility's ability to work on a cycle-by-cycle basis to fre- quencies below a few kilohertz [33, 34], disabling the ratchet mechanism for higher frequencies. It is noteworthy that the ion channels of apical OHCs differ significantly from those of basal OHCs [35] and that apical OHCs are considerably longer than basal ones and can accordingly produce greater length changes [25]. 3 0123Coefficientα0123Coefficientα0123Coefficientα102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)102103104Resonantfrequency(Hz)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)10−1100101102103104Sensitivity(nm·Pa−1)-6-4-2000.20.40.60.81Phase(cycles)RelativepositionABCDPeakedtravelingwaveRatchetmechanismIIIIIIα∗fHfLf1f2HBBMHBBM×90,000×10-6-4-2000.20.40.60.81Phase(cycles)RelativepositionABCDPeakedtravelingwaveRatchetmechanismIIIIIIα∗fHfLf1f2HBBMHBBM×90,000×10-6-4-2000.20.40.60.81Phase(cycles)RelativepositionABCDPeakedtravelingwaveRatchetmechanismIIIIIIα∗fHfLf1f2HBBMHBBM×90,000×10-6-4-2000.20.40.60.81Phase(cycles)RelativepositionABCDPeakedtravelingwaveRatchetmechanismIIIIIIα∗fHfLf1f2HBBMHBBM×90,000×10 A Unified Model for Cochlear Mechanics. We have quantified these considerations in a one-dimensional model of the cochlea (Supporting Information). Studies of threshold tuning curves of auditory-nerve fibers in the chinchilla have delineated three distinct regimes that are separated by two crossover frequen- cies, a high frequency fH ≈ 5 kHz and a low frequency fL ≈ 1.5 kHz (Fig. 3) [9]. In our model for the basal regime I, above fH, electromotility is negligible owing to the mem- brane time constant; each segment of the basilar membrane is tuned to its characteristic frequency. In the transitional regime II, between fH and fL, unidirectional coupling pro- vided by electromotility occurs but every segment of the basi- lar membrane still resonates at its characteristic frequency. In the apical regime III, below fL, electromotility combines with active hair-bundle motility to implement the ratchet mecha- nism for amplification. The resonant frequency of the basilar membrane is therefore of minor importance and remains ap- proximately constant at a value significantly above the range of characteristic frequencies. In accordance with the theory of critical-layer absorp- tion [1], in regime I the traveling wave produced by a pure- tone stimulus peaks at the characteristic place and sharply decreases beyond that point (Fig. 3C). The characteristic behavior of critical-layer absorption -- decrease of the travel- ing wave's speed and wavelength near the characteristic place -- appears in the phase behavior as an increase of the slope Fig. 4. Threshold tuning curves of auditory-nerve fibers. (A) The tuning curve at each position along the cochlea has a characteristic frequency f0 corresponding to the resonant frequency of the hair-bundle complex. When tuning curves are rescaled such that the frequency is measured in octaves relative to the characteristic frequency f0 and the threshold is measured relative to that at the characteristic frequency, char- acteristic shape changes are seen to occur between curves of different characteristic (B) Tuning curves for high characteristic frequencies, above fH, fall frequencies. onto a universal curve that exhibits the strongly asymmetric form and high-frequency cutoff characteristic of a peaked traveling wave. (C) As the characteristic frequency declines from fH to fL, the left limb falls (arrow), indicating the emerging influence of electromotility and the ratchet mechanism. (D) As the characteristic frequency diminishes below fL, the right limb falls steeply (arrow), pointing to the breakdown of the peaked-wave mechanism and the dominance of ratchet amplification. 4 (Fig. 3D). Active hair-bundle motility amplifies the peak dis- placement. A strong compressive nonlinearity arises because amplification by active hair-bundle motility both increases the hair-bundle and basilar-membrane motion per unit pressure difference and enhances the amplitude of the pressure wave itself (Fig. 3C). The combination of the two effects yields a compressive nonlinearity that extends over a significantly broader range of sound intensities than the nonlinearity in hair-bundle motion itself. In regime II electromotility influences the micromechan- ics and causes hair-bundle displacement to exceed basilar- membrane displacement. The theory of critical-layer ab- sorption still applies. The basilar-membrane tuning curves, phase behavior, and nonlinearity consequently parallel those in regime I. In regime III the ratchet mechanism leads to very dif- ferent behavior. Sound evokes a low-frequency pressure wave that traverses the entire cochlea, evoking only a small basilar-membrane displacement throughout, for the resonant frequency of the basilar membrane is higher everywhere (Fig. 3B). At the characteristic place, however, the hair bun- dles exhibit an independent resonance amplified through ac- tive hair-bundle motility and facilitated by the unidirectional coupling provided by electromotility. The hair-bundle dis- placement there can exceed the basilar-membrane response by orders of magnitude (Fig. 3C). In further contrast to the theory of critical-layer absorption, the traveling wave does not slow and its wavelength does not vanish at the characteristic place, for the basilar membrane does not resonate. This be- comes apparent in the behavior of the traveling wave's phase, whose slope remains low and nearly constant across the char- acteristic place (Fig. 3D). The phase behavior also confirms that the wave reaches the helicotrema. Because amplification through the ratchet mechanism does not act on the basilar membrane, the latter exhibits approximately linear behavior (Fig. 3C). The pressure wave is therefore unaffected by the active process. Hair-bundle displacement is amplified and ex- hibits a moderate compressive nonlinearity (Fig. 3C). Threshold Tuning Curves of Auditory-Nerve Fibers. Strong support for the proposed model comes from studies of thresh- old tuning curves for auditory-nerve fibers [7, 8, 9]. Recent measurements have shown characteristic shape changes oc- curring around the crossover frequencies fH and fL [9]. To compare these data to our model, we have computed tuning curves and -- despite their complexity and the simplicity of our model -- found striking agreement with the measurements (Fig. 4). High-frequency fibers, those tuned above fH, dis- play the strongly asymmetric shape characteristic of critical- layer absorption. Moreover, they fall onto a universal curve when the frequency and threshold are measured relative to the values at the characteristic frequency. This accords with ex- perimental findings and the scaling symmetry of the peaked traveling-wave mechanism [4, 5, 9]. For intermediate char- acteristic frequencies between fL and fH, the scaling law is violated as the ratchet mechanism starts to influence the mi- cromechanics, causing the left limb of the threshold tuning curves to fall as the characteristic frequency decreases [9]. A second violation of the scaling arises for characteristic frequen- cies below fL, for which the right limb falls. Also observed experimentally [9], this second violation of scaling indicates a breakdown of critical-layer absorption, which predicts a steep increase for threshold tuning curves at frequencies above the characteristic frequency. 0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfL0102030405060708090100100010000Threshold(dBSPL)Frequency(Hz)AfHfLf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequencyf04f02f02f0Frequency0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH0102030405060f04f02f02f0Rel.thresh.(dBSPL)FrequencyDf0≤fLCfL≤f0≤fHBf0≥fH Discussion The classical theory of cochlear mechanics assumes that the basilar membrane resonates at a characteristic place for each frequency in the auditory range. The resulting mechanism of critical-layer absorption is characterized by (i) strongly asymmetric, scale-invariant tuning curves with steep high- frequency cutoffs, (ii) an increasing slope of the traveling wave's phase upon approaching the characteristic place, (iii) basilar-membrane displacement similar to hair-bundle dis- placement, and (iv) pronounced compressive nonlinearity at the characteristic frequency. This picture is consistent with di- verse experimental findings from the cochlear base, including direct measurements of basilar-membrane motion and studies of threshold tuning curves for auditory-nerve fibers. However, as summarized in the Introduction, experimental results from the cochlea's apex are in qualitative disagreement with the first, second, and fourth characteristics. Here we have proposed a concept for low-frequency hear- ing that employs a ratchet mechanism involving the interplay of active hair-bundle motility and electromotility. Whereas sound-evoked forces displace the basilar membrane and the hair bundles, the active forces within the hair bundle can de- couple from the basilar membrane through appropriate elon- gation and contraction of OHCs. This mechanism allows hair bundles to resonate independently of the basilar membrane and thus explains how the ear's hearing range can extend to values well below the resonant frequency of the basilar membrane. We have shown that amplification by the ratchet mechanism leads to characteristic behavior that is distinct from that associated with critical-layer absorption: it exhibits (i(cid:48)) approximately symmetric tuning curves that display no sharp high-frequency cutoff, (ii(cid:48)) a constant slope of the trav- eling wave's phase across the characteristic place, (iii(cid:48)) tuned hair-bundle displacement that exceeds the untuned basilar- membrane displacement by orders of magnitude at resonance, and (iv(cid:48)) approximately linear basilar-membrane displacement and only moderate compressive nonlinearity in hair-bundle motion at the characteristic frequency. Comparison to Experimental Results. The proposed model yields the classical theory of critical-layer absorption for high- frequency sounds (Figs. 1B, 3 and 4A). Experiments in the base have confirmed (i) the asymmetric form of tuning curves [13, 14], (ii) the increasing slope of the phase [36, 37], and (iv) the strong compressive nonlinearity [13, 37, 14]. Ow- ing to difficulties in accessing the motion of the organ of Corti at the base, the relation (iii) of hair-bundle motion to basilar- membrane motion has not yet been measured there. Because the membrane time constant of OHCs is ex- pected to limit the operation of electromotility to low frequen- cies [33, 34], we have chosen in our model to consider only ac- tive hair-bundle motility in the basal part of the cochlea. How- ever, studies in prestin knockout mice suggest that electro- motility is necessary for high-frequency amplification [38, 39]. Electromotility near the base may adjust the operating point of the hair-bundle motor on a time-scale slower than the pe- riod of oscillation or provide a fast electrical signal in the form of membrane-potential change upon mechanical stimulation. Our theory accounts for a number of unexplained results from the cochlear apex. Experiments have measured (i(cid:48)) approximately symmetric tuning curves without sharp high- frequency cutoffs of tectorial-membrane and therefore hair- bundle displacement [10, 15, 11]. In our model the approxi- mate symmetry arises naturally, for it reflects the resonance of the hair bundles, reticular lamina, and tectorial membrane in- dependently of resonance by the basilar membrane (Figs. 1C and 3C,D). Experiments have consistently reported (ii(cid:48)) a con- stant phase slope across the characteristic place [10, 40, 11]. This feature emerges in our model (Fig. 3D) because the basi- lar membrane is untuned at low frequencies and the pressure wave therefore reaches the helicotrema. There remains a con- troversy about the existence of nonlinearity at the cochlear apex (reviewed in Ref. [12]): although some investigators have found no nonlinearites [15], others have reported a small compressive [10, 16] or expansive nonlinearity [11]. However, all agree about (iv(cid:48)) the absence of a strong compressive non- linearity. The cochlear apex allows observation of different parts of the organ of Corti -- such as the tectorial membrane and Hensen's cells as well as the basilar membrane -- and thus per- mits tests of characteristic (iii(cid:48)) of our model. Experiments on a temporal-bone preparation [15, 41] have shown that vi- brations of the reticular lamina and tectorial membrane are orders of magnitude larger than the basilar-membrane dis- placement at the characteristic place. These findings are in perfect agreement with our theory. However, the in vitro ex- periments have been criticized for an uncertainty about which constituents of the organ of Corti were measured [12] and are in contradiction with studies using other preparations [10, 16]. Definite conclusions must therefore be postponed until further experimental results become available. Perhaps the most reliable comparison of our model to experimental measurements comes through tuning curves of auditory-nerve fibers, which yield information about cochlear mechanics that is least disturbed by experimental intervention and represent responses from the whole length of the cochlea. These tuning curves exhibit progressive shape changes from (i) the strongly asymmetric form typical of high-frequency fibers to (i(cid:48)) the nearly symmetric form found for low-frequency fibers [7, 8, 9]. Our modeled responses of tuning curves of auditory-nerve fibers exhibit the same shape changes (Fig. 4), thus providing a conceptual explanation for the three distinct cochlear regimes. Experimentally Testable Predictions. Our theory yields ex- perimentally testable predictions based on the characteris- tics (i(cid:48))-(iv(cid:48)). As discussed above, (i(cid:48)) the approximately symmetric tuning curves as well as (ii(cid:48)) the constant phase slope have already been observed. In contrast, characteris- tic (iii(cid:48)), which predicts hair-bundle motion that exceeds the basilar-membrane motion at the characteristic frequency by orders of magnitude, remains controversial. Future exper- iments are therefore required to test this prediction. Such studies should also determine whether, as our theory implies, the complex formed by the hair bundles, reticular lamina, and tectorial membrane exhibits a resonance independent of the basilar membrane, whose response is untuned. A further test is feasible through experimental studies of the nonlin- ear behavior at the apex, for which our theory predicts (iv(cid:48)) a moderate compressive nonlinearity in hair-bundle motion, and therefore in tectorial-membrane motion, but no more than a weak nonlinearity in the basilar membrane's response. Fi- nally, and beyond the theory presented in this article, the ratchet mechanism should allow for traveling waves in the tectorial membrane [42] that are unidirectionally coupled to basilar-membrane waves. The Ratchet Principle. The ratchet mechanism constitutes a general design principle for mechanical amplification whereby the output does not feed back onto the input. In this way it represents a mechanical analogue of the operational amplifier from electrical engineering [28]. Although the current technol- ogy of signal detectors such as microphones relies on electrical amplification, mechanical amplification could improve signal- 5 to-noise ratios and thereby greatly advance sensitivity and detection of weak signals. The ratchet mechanism thus opens a path for implementing controlled mechanical amplification in engineering applications. Materials and Methods Hydrodynamics. By combining continuity equations and fluid-momentum equa- tions, we describe the cochlea's hydrodynamics with the partial differential equation ρ∂2 t XBM + Λ∂tXBM = 1 2L2 ∂r (h∂rp) . [ 4 ] Here ρ denotes the density of liquid in the cochlea, L the length of the cochlea, h the height of the scalae, and p and XBM respectively the pressure across and the displacement of the basilar membrane at position r and time t. The coefficient Λ accounts for friction due to fluid motion. Position is measured in units of the cochlear length, such that r = 0 corresponds to the basal and r = 1 to the apical end. The pressure translates into an external force acting on the basilar membrane and yields a displacement that we compute employing the model of Fig. 2B (Supporting Information). Temporal Fourier transformation yields an ordinary differential equa- tion that we solve numerically with the shooting method in Mathematica 7 (Wolfram Research). We apply two boundary conditions. First, p = p0 at r = 0: a sound- evoked pressure p0 acts at the stapes. And second, p = 0 at r = 1: because the two scalae communicate at the helicotrema, the pressure difference between them vanishes at the apical end of the cochlea. Tuning Curves of Auditory-Nerve Fibers. Tuning curves of auditory-nerve fibers are computed by assuming that the hearing threshold corresponds to a root-mean- square deflection of the hair bundles of IHCs by 0.3 nm. These bundles are thought to be coupled by fluid motion to the shearing between the reticular lamina and tecto- rial membrane, and thus to the displacement of the hair bundles of OHCs (Supporting Information). ACKNOWLEDGMENTS. We thank S. Leibler for discussion, the two reviewers for valuable suggestions, and the members of our research group for comments on the manuscript. This research was supported by grant DC00241 from the National In- stitutes of Health and by a fellowship to T. R. from the Alexander von Humboldt Foundation. A. J. H. is an Investigator of Howard Hughes Medical Institute. 26. X. Dong, M. Ospeck, and K. H. Iwasa. Piezoelectric reciprocal relationship of the membrane motor in the cochlear outer hair cell. Biophys. J., 82:1254 -- 1259, 2002. 27. B. N. Evans and P. Dallos. Stereocilia displacement induced somatic motility of cochlear outer hair cells. Proc. Natl. Acad. Sci. U.S.A., 90:8347 -- 8351, 1993. 28. P. Horowitz and W. Hill. The Art of Electronics. Cambridge University Press, second 29. edition, 1989. I. Sziklai. The significance of the calcium signal in the outer hair cells and its possible role in tinnitus of cochlear origin. Eur. Arch. Otorhinolaryngol., 261:517 -- 525, 2004. 30. J. J. Zwislocki and E. J. Kletsky. Tectorial membrane: a possible effect on frequency analysis in the cochlea. Science, 204:639 -- 641, 1979. 31. A. W. Gummer, W. Hemmert, and H. P. Zenner. Resonant tectorial membrane motion in the inner ear: its crucial role in frequency tuning. Proc. Natl. Acad. Sci. U.S.A., 93:8727 -- 8732, 1996. 32. H. Cai, B. Shoelson, and R. S. Chadwick. Evidence of tectorial membrane radial motion in a propagating mode of a complex cochlear model. Proc. Natl. Acad. Sci. U.S.A., 101:6243 -- 6248, 2004. 33. J. Santos-Sacchi. On the frequency limit and phase of outer hair cell motility: effects of the membrane filter. J. Neurosci., 12:1906 -- 1916, 1992. 34. G. D. Housley and J. F. Ashmore. Ionic currents of outer hair cells isolated from the guinea-pig cochlea. J. Physiol., 448:73 -- 98, 1992. 35. J. Engel, C. Braig, L. Ruttiger, S. Kuhn, U. Zimmermann, N. Blin, M. Suasbier, H. Kalbacher, S. Munkner, K. Rohbock, P. Ruth, H. Wintner, and M. Knipper. Two classes of outer hair cells along the tonotopic axis of the cochlea. Neurosci., 143:837 -- 849, 2006. 36. W. S. Rhode. Observations of the vibration of the basilar membrane in squirrel mon- keys using the Mossbauer technique. J. Acoust. Soc. Am., 49:1218 -- 1231, 1971. 37. A. L. Nuttall and D. F. Dolan. Steady-state sinusoidal velocity responses of the basilar membrane in guinea pig. J. Acoust. Soc. Am., 99:1556 -- 1565, 1996. 38. M. M. Mellado Lagarde, M. Drexl, V. A. Lukashkina, A. N. Lukashkin, and I. J. Rus- sell. Outer hair cell somatic, not hair bundle, motility is the basis of the cochlear amplifier. Nat. Neurosci., 11:746 -- 748, 2008. 39. P. Dallos, X. Wu, M. A. Cheatham, J. Gao, J. Zheng, C. T. Anderson, S. Jia, X. Wang, W. H.Y. Cheng, S. Sengupta, D. Z. Z. He, and J. Zuo. Prestin-based outer hair cell motility is necessary for mammalian cochlear amplification. Neuron, 58:333 -- 339, 2008. 40. W. S. Rhode and N. P. Cooper. Fast traveling waves, slow traveling waves and their interactions in experimental studies of apical cochlear mechanics. Auditory Neurosci., 2:289 -- 199, 1996. International Team for Ear Research. Cellular vibration and motility in the organ of Corti. Acta Oto-Laryngol. Suppl., 467:1 -- 279, 1989. 41. 42. R. Ghaffari, A. J. Aranyosi, and D. M. Freeman. Longitudinally propagating trav- eling waves of the mammalian tectorial membrane. Proc. Natl. Acad. Sci. U.S.A., 104:16510 -- 16515, 2007. 1. J. Lighthill. Energy flow in the cochlea. J. Fluid Mech., 106:149 -- 213, 1981. 2. F. Mammano and R. Nobili. Biophysics of the cochlea: Linear approximation. J. Acoust. Soc. Am., 93:3320 -- 3332, 1993. 3. C. A. Shera, A. Tubis, and C. L. Talmadge. Do forward- and backward-traveling waves occur within the cochlea? Countering the critique of Nobili et al. JARO, 5:349 -- 359, 2004. 4. W. M. Siebert. Stimulus transformations in the peripheral auditory system. In P. A. Kolers and M. Eden, editors, Recognizing Patterns, pages 104 -- 133. MIT, Cambridge, 1968. 5. G. Zweig. Basilar membrane motion. Cold Spring Harb Symp Quant Biol, 40:619 -- 633, 1976. 6. R. C. Naidu and D. C. Mountain. Measurements of the stiffness map challenge a basic tenet of cochlear theories. Hear. Res., 124:124 -- 131, 1998. 7. N. Y. S. Kiang, M. C. Liberman, W. F. Sewell, and J. J. Guinan. Single unit clues to cochlear mechanics. Hear. Res., 22:171 -- 182, 1986. 8. M. Ulfendahl. Mechanical responses of the mammalian cochlea. Progr. Neurobiol., 53:331 -- 380, 1997. 9. A. N. Temchin, N. C. Rich, and M. A. Ruggero. Threshold tuning curves of chinchilla auditory-nerve fibers. I. Dependence on characteristic frequency and relation to the magnitudes of cochlear vibrations. J. Neurophysiol., 100:2889 -- 2898, 2008. 10. N. P. Cooper and W. S. Rhode. Nonlinear mechanics at the apex of the guinea-pig cochlea. Hear. Res., 82:225 -- 243, 1995. 11. C. Zinn, H. Maier, H. P. Zenner, and A. W. Gummer. Evidence for active, nonlin- ear, negative feedback in the vibration response of the apical region of the in vivo quinea-pig cochlea. Hear. Res., 142:159 -- 183, 2000. 12. L. Robles and M. A. Ruggero. Mechanics of the mammalian cochlea. Physiol. Rev., 81:1305-1352, 2001. 13. N. P. Cooper and W. S. Rhode. Basilar membrane mechanics in the hook region of cat and guinea-pig cochlea: sharp tuning and nonlinearity in the absence of baseline position shift. Hear. Res., 63:191 -- 196, 1992. 14. M. A. Ruggero, N. C. Rich, A. Recio, S. S. Narayan, and L. Robles. Basilar-membrane J. Acoust. Soc. Am., responses to tones at the base of the chinchilla cochlea. 101:2151 -- 2163, 1997. 15. S. M. Khanna and L. F. Hao. Reticular lamina vibrations in the apical turn of a living guinea pig cochlea. Hear. Res., 132:15 -- 33, 1999. 16. W. S. Rhode and N. P. Cooper. Nonlinear mechanics in the apical turn of the chinchilla cochlea in vivo. Auditory Neurosci., 3:101 -- 121, 1996. 17. P. Martin and A. J. Hudspeth. Active hair-bundle movements can amplify a hair cell's response to oscillatory mechanical stimuli. Proc. Natl. Acad. Sci. U.S.A., 96:14306 -- 14311, 1999. 18. P. Martin, A. J. Hudspeth, and F. Julicher. Comparison of a hair bundle's spontaneous oscillations with its response to mechanical stimulation reveals the underlying active process. Proc. Natl. Acad. Sci. U.S.A, 98:14380 -- 14385, 2001. 19. H. J. Kennedy, A. C. Crawford, and R. Fettiplace. Force generation by mammalian hair bundles supports a role in cochlear amplification. Nature, 433:880 -- 883, 2005. 20. D. K. Chan and A. J. Hudspeth. Ca2+ current-driven nonlinear amplification by the mammalian cochlea in vitro. Nat. Neurosci., 8:149 -- 155, 2005. 21. B. Nadrowski, P. Martin, and F. Julicher. Active hair-bundle motility harnesses noise to operate near an optimum of mechanosensitivity. Proc. Natl. Acad. Sci. U.S.A., 101:12195 -- 12200, 2004. 22. A. J. Hudspeth. Making an effort to listen: Mechanical amplification in the ear. Neuron, 59:530 -- 545, 2009. 23. J. Zheng, W. Shen, D. Z. Z. He, K. B. Long, L. D. Madison, and P. Dallos. Prestin is the motor protein of cochlear outer hair cells. Nature, 405:149 -- 155, 2000. 24. P. Dallos, J. Zheng, and M. A. Cheatham. Prestin and the cochlear amplifier. J. Physiol., 576:37 -- 42, 2006. 25. J. Ashmore. Cochlear outer hair cell motility. Physiol. Rev., 88:173 -- 210, 2008. 6 SupportingInformationARatchetMechanismforAmplificationinLow-FrequencyMammalianHearingTobiasReichenbachandA.J.HudspethHowardHughesMedicalInstituteandLaboratoryofSensoryNeuroscience,TheRockefellerUniversity,NewYork,NewYork10065-6399,U.S.A.Herewefurtherdescribeourmodelingapproach.Weprovideadetaileddescriptionofthetwo-massmodelfortheorganofCortianddiscusstheenergybalance.Wethenturntothehydrodynamicsofthecochlea.Becauseourgoalistodemonstrateaprincipleratherthantodescribethecochleainfulldetail,wefocusonabasic,one-dimensionalmodel.WeconcludewithamodelfortheregulationoftheOHCs'membranepotentialwhichdemonstrateshowthecriticalmechanomotilitycoefficientα∗canresult.MicromechanicsoftheorganofCortiTodescribethemicromechanicsoftheorganofCortiweemployamodelwithtwodegreesoffreedom:themotionofthebasilarmembraneandthemotionofthehair-bundlecomplex(Figure2B).TheimpedanceZBMofthebasilarmembraneresultsfromamassmBM,viscousdampingλBM,andstiffnessKBM.Similarly,theimpedanceofthecomplexformedbythehairbundles,reticularlamina,andtectorialmembraneinvolvesmassmTM,viscousdampingλHB,andstiffnessKHB.Themotionsofthebasilarmembraneandofthehair-bundlecomplexarecoupledthroughtheimpedanceZCoftheorganofCorti,whichpossessesviscousandelasticcontributionsλCandKC.FurthercouplingarisesthroughthecombinedOHCsandDeiters'cells(Fig.S1).AnOHCcanbedescribedasapiezoelectricelement[1].ItslengthchangeδLdependsbothonachangeδVinthemembranepotentialaswellasontheappliedforceF.Consideringoscillatorymotionsatangularfrequencyω=2πfwecanwriteδL=fδLeiωt+c.c.,δV=fδVeiωt+c.c.,andF=eFeiωt+c.c.inwhichc.cdenotesthecomplexconjugate.ThelengthchangethenfollowsasfδL=−cfδV+c2eF(5)withcoefficientsc,c2.WedefineXEE=−cfδVasthepartofthelengthchangethatresultsfromvoltagechangesalone.Thelengthchangec2eFresultsfromtheOHC'sviscoelasticityZOHC=−iω−1c−12inserieswithXEE(Fig.S1).ItliesinserieswiththeDeiters'cellwhichcanbedescribedasanotherviscoleasticelementZDC.ThetwoviscoelasticelementsZOHCandZDCinseriescombineintotheviscoelasticelementZD=(Z−1OHC+Z−1DC)−1(6)1 .OHCRLDCBMZDCZOHCXHBδLXBMZCXEEFigureS1:Couplingbetweenthereticularlamina(RL)andthebasilarmembrane(BM).ThecouplingarisesfromOHCsinserieswithDeiters'cells(DC)andanimpedanceZCfromtheremainingorganofCorti.SeethetextforadescriptionofthecoupledOHCsandDeiters'cells.suchthattheschematicofFig.2Bresults.WedenotebyKDtheelasticcomponentandbyλDtheviscouspartoftheimpedanceZD.Theequationsofmotionforhair-bundleandbasilar-membranedisplacementXHBandXBMthenreadmTM∂2tXHB+λHB∂tXHB+KHBXHB+(KD+λD∂t)(XHB−XEE−XBM)+(KC+λC∂t)(XHB−XBM)=Fint,mBM∂2tXBM+λBM∂tXBM+KBMXBM−(KD+λD∂t)(XHB−XEE−XBM)−(KC+λC∂t)(XHB−XBM)=Fext.(7)AsoundstimulusatangularfrequencyωprovidesanexternalforceFextonthebasilarmembranewhosedependenceontimetmaybewrittenasFext(t)=Fexteiωt+c.c.withtheFouriercoefficientFext.Theevokedoscillationsofthebasilarmembrane,XBM,andofthehair-bundlecomplex,XHB,occurdominantlyatthesamefrequencyf;additionalfrequenciesmayarisefromnonlineareffectsowingtointernalforcesinthehairbundle.WeconsiderthecorrespondingFouriercoefficientsXBMandXHB.TheelectricallyevokedlengthchangeXEEofOHCsdependslinearlyonthehair-bundledisplacement,forthemembrane-potentialchangedependslinearlyonhair-bundlemotion(seeSupplementarySectionontheelectricalregulationoftheOHCmembranepotential):XEE=−αXBMwithacomplex,frequency-dependentmechanomotilitycoefficientα.SupplementaryEquation(7)thenyieldsEquations(1)and(2)withZHB=iωmTM+λHB−iω−1KHB,ZBM=iωmBM+λBM−iω−1KBM,ZC=λC−iω−1KC,ZD=λD−iω−1KD.(8)Theforceswithinthehairbundledependonitsdisplacement;theireffectistocounterviscousdampingaswellastoinfluencetheresonantfrequencyofthehair-bundlecomplex.Wemaydecomposetheseforcesintolinearand2 nonlinearparts:Fint=iωZactHBXHB+Fnonlinint(XHB).(9)Herethelineartermdominatesthedynamicsforsmalldisplacementsensuingfromweaksoundstimulianddescribesthefullyactivescenario.Thenonlineartermconnectstheactivetothepassivecasethatarisesforstrongsoundstimuli.Atthecriticalvalueα∗,thedisplacementsaregivenbyEquations(3),whichcanberewrittenusingSupple-mentaryEquation(9)asXHB=1iω(ZHB−ZactHB)Fnonlinint(XHB)+ZD+ZCiω(ZHB−ZactHB)(ZBM+ZD+ZC)Fext,XBM=1iω(ZBM+ZD+ZC)Fext.(10)Resonanceinhair-bundlemotionoccursatthefrequencyforwhichZHB−ZactHB=0.Activehair-bundlemotilitymakesthissituationpossible.Asdiscussedinthemaintext,thisresonanceconditionisindependentoftheprop-ertiesofthebasilarmembrane.CochlearhydrodynamicsandtravelingwavesWenowincorporatethedescriptionforthemicromechanicsoftheorganofCortiintoamodelforthewholecochlea.Inthesimplestform,thecochleaisconsideredasone-dimensional,exhibitingaslowpressurewave.Combiningcontinuityequationsandfluid-momentumequations,thispressurewaveobeysthepartialdifferentialequationρ∂2tXBM+Λ∂tXBM=12L2∂r(h∂rp).(11)Hereρdenotesthedensityofliquidinthecochlea,Lthelengthofthecochlea,htheheightofthescalae,andpandXBMrespectivelythepressureacrossandthedisplacementofthebasilarmembraneatpositionrandtimet.ThecoefficientΛaccountsforfrictionduetofluidmotion.Positionismeasuredinunitsofthecochlearlength,suchthatr=0correspondstothebasalandr=1totheapicalend.Weapplytwoboundaryconditions.First,p=p0atr=0:asound-evokedpressurep0actsatthestapes.Andsecond,p=0atr=1:becausethetwoscalaecommunicateatthehelicotrema,thepressuredifferencebetweenthemvanishesattheapicalendofthecochlea.ConsideringtheFouriercoefficientsofangularfrequencyω,weobtain−ω2ρXBM+iωΛXBM=12L2∂r(h∂rp).(12)Thebasilar-membranedisplacementXBMandpressurepareconnectedthroughEquation(1),withp=Fext/ABM.ABMdenotestheareaofatransversestripofthebasilarmembranethathasthewidthofonecell,about8µm;allparametersemployedinthetwo-massmodelrefertoatransversesegmentofthecochlearpartitionofthatwidth.WesolveSupplementaryEquation(12)numericallywiththeshootingmethodinMathematica7(WolframResearch).3 Theheighthofthescalaevariesfromabout1mmatthebasetoabout200µmattheapex.Measurementsofthefluidpressureinthebasalregionindicate,however,thatthepenetrationdepthofthewaveissignificantlysmaller[2].Inourone-dimensionalsimulations,wethereforeuseaneffectiveheightof100µm.Concerningthemapofcharacteristicfrequenciesf0(r)alongthecochlea,weconsiderthechinchilla'shearingrangefromamaximalfrequencyofabout30kHzattheextremebasetoaminimumofabout50Hzattheextremeapex[3].WeassumethattheresonantfrequencyofthebasilarmembranecannotcoverthiswholerangeandisinsteadgivenbyfBM(r)=qf2BM,apex+[1−(fBM,apex)/fmax)2]f0(r)2.(13)Herefmax=30kHzdenotesthehighestcharacteristicfrequencyandfBM,apex=1kHzthelowestresonantfrequencythatthebasilarmembraneexhibits.fBMcoincideswithf0forhighfrequenciesbutdeviatesforfrequenciesbelowfLandapproachesfBM,apexintheapicalregionofthecochlea(Figure3B).TheresonantfrequencyfBMofthebasilarmembraneissetbyitsmassandstiffnessaccordingtofBM=(2π)−1pKBM/mBM.WeconsiderKBMtovaryproportionaltofBMandmBMtovaryinversely:KBM(r)=KmaxBMfBM(r)/fmaxandmBM(r)=KBM(r)/[2πfBM(r)]2,withKmaxBM=2N·m−1.ThemassmTMofthetectorialmembraneisassumedtofollowthesamespatialvariationasthebasilarmembranemass,albeitwithasmallervalue:mTM=mBM/5.Theincreaseinthebasilar-membraneandtectorial-membranemassfrombasetoapexpresumablyreflectstheincreasingsizeoftheorganofCortitowardstheapex.Forthevariationofthemechanomotilitycoefficientαalongthecochlea,anditsdependenceonthestimulusfrequencyf,wemaketheansatzα(r,f)=f4cf4c+f0(r)4×[δf(r)f0]2[f2−f0(r)2]2+[δff0(r)]2×α∗.(14)Thefirstfactoraccountsforthehigh-frequencycutoffabovewhichthemembranepotential,andthuselectromotility,cannolongerfollowhair-bundledisplacementonacycle-by-cyclebasis;weusefc=4kHz.Toyieldtheratchetmechanismatlowerfrequencies,wechooseαtobeα∗atthenaturalfrequency.Thesecondfactordescribeshowtheresponsechangeswhenthefrequencyfofstimulationdeviatesfromthenaturalfrequencyf0;weassumethatαisclosetoα∗aslongasbothfrequenciesaresimilar,butotherwisedeclinesinmagnitude.Thewidthofthecorrespondingcurveisdeterminedbytheparameterδ,whichwesetatδ=2.Themechanomotilitycoefficientα(r,f=f0)isshowninFigure3A.Thelinearpartoftheactivehair-bundleforcecountersviscousdampingandprovidesavanishingimpedanceatthenaturalfrequency.UsingEquations(1),(2),andSupplementaryEquation(9),thistranslatesintoZHB−ZactHB(cid:12)(cid:12)(cid:12)f=f0=−ZBM[(1+α)ZD+ZC]ZBM+ZC+ZD(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)f=f0.(15)WeassumethattheimaginarypartofZHBistunedtothenaturalfrequencyf0suchthattheactivecontributionpossessesonlyarealpartcorrespondingtonegativedamping,ZactHB≡λactHB.Notethatintheratchetmechanism,whenα=α∗,theright-handsideofSupplementaryEquation(15)vanishes,soactivehair-bundleforcesneedtoovercomeonlydampingassociatedwithhair-bundlemotion.4 ParameterDescriptionValueReferenceLlengthofthecochlea20mm[4]Λfluidfriction5×105N·s·m−4wBM(r)widthofthebasilarmembrane(50+200r)µm[5]ABM(r)areaofthebasilar-membranestripwBM(r)×8µmλHBfrictionofthehair-bundlecomplex1µN·s·m−1[6]KIHBstiffnessofthehairbundlesofIHCs2mN·m−1[7]λIHBfrictionofthehairbundlesofIHCs1µN·s·m−1[6]λBM(r)frictionofthebasilarmembrane0.03×wBM(r)N·s·m−2KCstiffnessoftheorganofCorti10mN·m−1λCfrictionoftheorganofCorti10µN·s·m−1KDstiffnessofOHCsandDeiters'cells20mN·m−1[8]λDfrictionofOHCsandDeiters'cells10µN·s·m−1S1:Parametersusedinthecochlearmodel.Whenαandactivehair-bundleforcesobeySupplementaryEquations(14)and(15)andnonlinearitiesfromthehair-bundleforcesareignored,thehair-bundledisplacementdivergesatresonance.Intheactualcochlea,noiseaswellasnonlinearitiessuperveneandleadtoalargebutfinitedisplacement.Inoursimulations,weincorporatethiseffectbyconsideringvaluesforαandZactHBthatdeviateby1%fromtheiridealvaluesgivenbySupplementaryEquations(14)and(15).Tuningcurvesofauditory-nervefibersTuningcurvesofauditory-nervefibersarecomputedbyassumingthatthehearingthresholdcorrespondstoaroot-mean-squaredeflectionofthehairbundlesofIHCsby0.3nm.Thesebundlesarethoughttobecoupledbyfluidmotiontotheshearingbetweenthereticularlaminaandtectorialmembrane,andthustothedisplacementofthehairbundlesofOHCs.AssumingthatthedisplacementXIHBofthehairbundlesofinnerhaircellsisdependentonfrictionλIHBandelasticityKIHB,itcouplestotheOHCs'bundlemotionthroughtherelationλIHB∂t(XIHB−XHB)+KIHBXIHB=0.(16)UponFouriertransformation,weobtainXIHB=iωλIHBiωλIHB+KIHBXHB.(17)Forhighfrequencies,viscouscouplingisstrongandthehair-bundledisplacementsofIHCsandOHCscoincide.Forlowfrequencies,thecouplingdecreasesandthehair-bundlemotionofIHCsremainssmallerthanthatofOHCs.Thiseffectunderliestherisingthresholdsatlowfrequenciesseenontheleftlimbsofhigh-frequencytuningcurves(Figure5)aswellasinexperimentalmeasurements.Thevalueof0dBSPLisdefinedbyaroot-mean-squaresound-pressurestimulusof20µPa.Becausethis5 ABNa+Ca2+,K+K+XHBVVERaCbVPRbFigureS2:RegulationoftheOHCmembranepotential.(A)Hair-bundledeflectionopensionchannels,K+andCa2+flowintothecell.ThecellmembranecontainsvariousadditionalionchannelsthatareregulatedbythemembranepotentialandtheCa2+concentration.Inourmodel,weconsidervoltage-regulatedoutwardK+channelsandinwardNa+channels.(B)AnequivalentcircuitofanOHCwithavariableresistanceRacontrolledbyhair-bundledisplacement,amembranecapacitanceCb,andabasoleteralmembraneconductanceRb.externalpressureisenhancedbythemiddleearbeforeactingonthestapes,weassumeanincreasebyafactorof20(26dB)[9].Iftheyhavenotbeendetailedinthetextabove,theparametersofthetwo-massmodelaswellastheone-dimensionaldescriptionsofthecochleaandhairbundledynamicsofIHCsaregiveninTableS1.ElectricalregulationoftheOHCmembranepotentialDeflectionXHBoftheOHCshairbundlesevokesachangeδV≡V−V0inthemembranepotentialfromitsrestingvalueV0,whichleadstoalengthchangeXEE=−cδV(seeSupplementaryEquation5andbelow)withacoefficientcofabout20µm·V−1[1].Theratchetmechanismreliesonacriticalvalueα∗=−1−ZC/ZDoftheratioα=−XEE/XHBbetweenhair-bundledisplacementandcelllengthchange.WhileitsexactvaluedependsonthecouplingimpedancesZCandZD,therealpartofα∗isnegative.AtypicalsituationZC=ZDleadstothecriticalvalueα∗=−2.Hereweshowhowsuchavalueofα∗canemergewithinabasicmodelfortheregulationoftheOHCsmembranepotential.Themodelreliesonvoltage-regulatedoutwardK+channels[10]thatcounteractthecurrentthroughthemembranecapacitance,aswellasavoltage-regulatedinwardsodiumcurrent[11;12]thatcanleadtohyperpolarizationuponpositivehair-bundledeflection.TheapicalsurfaceofanOHC,includingthehairbundle,isbathedinendolymphatapotentialVEofabout80mV(FigureS2).ThebasolateralpartissurroundedbyperilymphatgroundpotentialVP.Theelectricalbehaviorcanberepresentedbyasimplecircuitdescribedbyanequationforcurrentconservation:IRa=IRb+ICb.(18)6 00.511.522.533.5101102103104060120180240300360MechanomotilitygainαPhaseofα(◦)Frequency(Hz)00.511.522.533.5101102103104060120180240300360MechanomotilitygainαPhaseofα(◦)Frequency(Hz)FigureS3:Themodeledmechanomotilitycoefficientα.Weshowthedependenceofthemagnitude(black)andphase(red)ofαonthesoundfrequency.Forf0=200Hzα=−2emergeswhichrepresentsatypicalvalueforthecriticalα∗.Themagnitudeofαisstronglyattenuatedforhigherfrequencies.ByvaryingthevoltagedependencyofthebasalK+andNa+channelsthecriticalvalueα∗canemergeatdifferentfrequenciesf0belowafewkilohertz.HereIRadenotesthecurrentthroughtheresistanceRaandIRbandICbdenotethecurrentsthroughRbandCb,respectively.WeassumethatRaresultsfromtheconductivitygaofthemechanotransductionchannelsinthehairbundle;thecurrentismainlycarriedbyK+ions[13].Theconductanceofthebasolateralcellmembranestemsfromvoltage-regulatedK+channels(conductancegKb)aswellasvoltage-regulatedNa+-channels[11;12](conductancegNab).Weobtainga(VE−V)=Cb∂tV+gKb(V−VP)−gNab(VNa−V)(19)inwhichVNadenotestheNa+reversalpotentialandCbthecellmembrane'scapacitance.ThepotentialV0insidethecellatvanishinghair-bundledisplacementXHB=0followsasV0=g(0)aVE+gNa(0)bVNa+gK(0)bVPg(0)a+gNa(0)b+gK(0)b(20)inwhichthesuperscript(0)denotestherespectiveconductancesatvanishinghair-bundledisplacement.Hair-bundledisplacementXHBchangestheconductancegaofthecell'sapicalmembrane.Tolinearoderweobtainga=g(0)a+∂ga∂XHB(cid:12)(cid:12)(cid:12)(cid:12)(0)XHB.(21)ThealteredapicalconductanceevokesachangesδV=V−V0inthemembranepotentialwhichinfluencesthebasalK+andNa+conductances.TheK+channelsarevoltage-regulatedatanactivationtimeτ.ThedynamicsofthechangeδgKb=gKb−gK(0)bcanbedescribedbyτ∂tδgKb=∂gKb∂V(cid:12)(cid:12)(cid:12)(cid:12)(0)δV−δgKb(22)7 ParameterDescriptionValueReferenceVEpotentialoftheendolymph80mV[14]VPpotentialoftheperilymph−100mV[14]VNaNa+reversalpotential22mV[11]gK+aapicalconductivityofK+channels5nS[15]gKbbasoleteralK+conductivity10nS[10]Cbmembranecapacitance40pF[10]∂gK+a/∂XHBdependenceofapicalK+conductivityonXHB5×10−3S·m−1[16]τactivationtimeofvoltage-regulatedK+channels1ms[10]S2:ParametersformodelingtheregulationoftheOHCsmembranepotential.whichresultsineδgKb=11+iωτ∂gKb∂V(cid:12)(cid:12)(cid:12)(cid:12)(0)fδV(23)foroscillatorystimuliXHB=XHBeiωt+c.c.atangularfrequencyω.TheNa+channelsarerapidlyactivated;tolinearordertheirconductancefollowsasgNab=gNa(0)b+∂gNab∂V(cid:12)(cid:12)(cid:12)(cid:12)(0)δV.(24)LinearizationofSupplementaryEquation(19)yields"iωCb+ga+gKb+gNab+1−iωτ1+ω2τ2∂gKb∂V(cid:12)(cid:12)(cid:12)(cid:12)(0)(V−VP)−∂gNab∂V(cid:12)(cid:12)(cid:12)(cid:12)(0)(VNa−V)#fδV=∂ga∂XHB(cid:12)(cid:12)(cid:12)(cid:12)(0)XHB.(25)Themechanomotilitycoefficientαfollowsasα=cfδV/XHB.FromSupplementaryEquation(25)wenotethatdelayedvoltage-regulatedK+channelscanreducetheeffectofthemembranecapacitance.AlargeenoughvoltagedependenceoftheinwardNa+channelscanthenyieldhyperpolarizationofthecelluponpositivedeflectionofthehairbundle,asrequestedforthecriticalvalueα∗.FigureS3showstheresultingmagnitudeandphaseofαfortypicalparametersforapicalOHCs(TableS2)independenceonthefrequencyofhair-bundlestimulation.Thevalues∂gKb/∂V=1.36µS·V−1and∂gNab/∂V=1.85µS·V−1havebeenchosentoyieldthecriticalvalueα∗=−2atthefrequencyf0=200Hzandlieintherangeofmeasuredvalues[10;11].8 References[1]X.Dong,M.Ospeck,andK.H.Iwasa.Piezoelectricreciprocalrelationshipofthemembranemotorinthecochlearouterhaircell.Biophys.J.,82:1254 -- 1259,2002.[2]E.Olson.Directmeasurementofintra-cochlearpressurewaves.Nature,402:526 -- 529,1999.[3]R.S.HeffnerandH.E.Heffner.Behavioralhearingrangeofthechinchilla.Hear.Res.,52:13,1991.[4]P.A.SantiandD.C.Muchow.Morphometryofthechinchillaorganofcortiandstriavascularis.J.Histochem.andCytochem.,27:1539 -- 1542,1979.[5]L.I.ProkofjevaandA.G.Chernyi.Quantitativecharacteristicsofthebasilarmembraneinthemammaliancochlea.NauchnyeDokiVyssShkolyBiolNauki.,11:44 -- 50,1986.[6]P.Martin,A.J.Hudspeth,andF.Julicher.Comparisonofahairbundle'sspontaneousoscillationswithitsresponsetomechanicalstimulationrevealstheunderlyingactiveprocess.Proc.Natl.Acad.Sci.U.S.A,98:14380 -- 14385,2001.[7]D.K.ChanandA.J.Hudspeth.MechanicalresponsesoftheorganofCortitoacousticandelectricalstimulationinvitro.Biophys.J.,89:4382 -- 4395,2005.[8]J.Ashmore.Cochlearouterhaircellmotility.Physiol.Rev.,88:173 -- 210,2008.[9]S.Puria,W.T.Peake,andJ.J.Rosowski.Sound-pressuremeasurementsinthecochlearvestibuleofhuman-cadaverears.J.Acoust.Soc.Am.,101:2754 -- 2770,1997.[10]F.MammanoandJ.F.Ashmore.Differentialexpressionofouterhaircellpotassiumcurrentsintheisolatedcochleaoftheguinea-pig.J.Physiol.,469:639 -- 646,1996.[11]C.M.Witt,H.Y.Hu,W.E.Brownell,andD.Bertrand.Physiologicallysilentsodiumchannelsinmammalianouterhaircells.J.Neurophysiol.,72:1037 -- 1040,1994.[12]H.P.ZennerD.Oliver,P.PlinkertandJ.P.Ruppersberg.Sodiumcurrentexpressionduringpostnataldevelopmentofratouterhaircells.PflugersArch.Eur.J.Physiol.,434:1432 -- 2013,1997.[13]A.J.Hudspeth.Howtheear'sworkswork.Nature,341:397 -- 404,1989.[14]P.Mistrik,C.Mullaley,F.Mammano,andJ.Ashmore.Three-dimensionalcurrentflowinalarge-scalemodelofthecochleaandthemechanismofamplificationofsound.J.Roy.Soc.Int.,6:279 -- 291,2009.[15]A.J.Ricci,A.C.Crawford,andR.Fettiplace.Tonotopicvariationintheconductanceofthehaircellmechanotransducerchannel.Neuron,40:983 -- 990,2003.[16]M.Beurg,M.G.Evans,C.M.Hackney,andR.Fettiplace.ALarge-ConductanceCalcium-SelectiveMechan-otransducerChannelinMammalianCochlearHairCells.J.Neurosci.,26:10992 -- 11000,2006.9
1602.08299
3
1602
2017-02-01T01:21:09
Theory of population coupling and applications to describe high order correlations in large populations of interacting neurons
[ "q-bio.NC", "physics.bio-ph" ]
To understand the collective spiking activity in neuronal populations, it is essential to reveal basic circuit variables responsible for these emergent functional states. Here, I develop a mean field theory for the population coupling recently proposed in the studies of visual cortex of mouse and monkey, relating the individual neuron activity to the population activity, and extend the original form to the second order, relating neuron-pair's activity to the population activity, to explain the high order correlations observed in the neural data. I test the computational framework on the salamander retinal data and the cortical spiking data of behaving rats. For the retinal data, the original form of population coupling and its advanced form can explain a significant fraction of two-cell correlations and three-cell correlations, respectively. For the cortical data, the performance becomes much better, and the second order population coupling reveals non-local effects in local cortical circuits.
q-bio.NC
q-bio
Theory of population coupling and applications to describe high order correlations in large populations of interacting neurons RIKEN Brain Science Institute, Wako-shi, Saitama 351-0198, Japan (Dated: September 24, 2018) Haiping Huang To understand the collective spiking activity in neuronal populations, it is essential to reveal basic circuit variables responsible for these emergent functional states. Here, I develop a mean field theory for the population coupling recently proposed in the studies of visual cortex of mouse and monkey, relating the individual neuron activity to the population activity, and extend the original form to the second order, relating neuron-pair's activity to the population activity, to explain the high order correlations observed in the neural data. I test the computational framework on the salamander retinal data and the cortical spiking data of behaving rats. For the retinal data, the original form of population coupling and its advanced form can explain a significant fraction of two-cell correlations and three-cell correlations, respectively. For the cortical data, the performance becomes much better, and the second order population coupling reveals non-local effects in local cortical circuits. PACS numbers: 87.19.L-, 02.50.Tt, 75.10.Nr I. INTRODUCTION To uncover the neural circuit mechanisms underlying animal behavior, e.g., working memory or decision making, is a fundamental issue in systems neuroscience [1, 2]. Recent developments in multi-neuron recording methods make simultaneous recording of neuronal population activity possible, which gives rise to the challenging computational tasks of finding basic circuit variables responsible for the observed collective behavior of neural populations [3]. The collective behavior arises from interactions among neurons, and forms the high dimensional neural code. To search for a low dimensional and yet neurobiologically plausible representation of the neural code, thus becomes a key step to understand how the collective states generate behavior and cognition. Correlations among neurons' spiking activities play a prominent role in deciphering the neural code [4]. Various models were proposed to understand the pairwise correlations in the population activity [5 -- 7]. Modeling these correlations sheds light on the functional organization of the nervous system [8]. However, as the population size grows, higher order correlations have to be taken into account for modeling synchronous spiking events, which are believed to be crucial for neural information transmission [9 -- 11]. In addition, the conclusion drawn from small size populations may not be correct for large size populations. Theoretical studies have already proved that high order interactions among neurons are necessary for generating widespread population activity [12, 13]. However, introduction of high order multi-neuron couplings always suffers from a combinatorial explosion of model parameters to be estimated from the finite neural spike train data. To account for high order correlations, various models with different levels of approximation were proposed, for example, the reliable interaction model [14] with the main caveat that the rare patterns are discarded during inference of the coupling terms, the dichotomized Gaussian model [7, 15] in which correlations among neurons are caused by common Gaussian inputs to threshold neurons, the K-pairwise model [16, 17] in which an effective potential related to the synchronous firing of K neurons was introduced, yet hard to be interpreted in terms of functional connectivity, and the restricted Boltzmann machine [18] where hidden units were shown to be capable of capturing high order dependences but their number should be predefined and difficult to infer from the data [19]. One can also take into account part of the statistical features of the population activity (e.g., simultaneous silent neural pattern) and assume homogeneity for high order interactions among neurons due to the population size limitation [20]. In this paper, I provide a low dimensional neurobiological model for describing the high order correlations and extracting useful information about neural functional organization and population coding. In this study, I interpret correlations in terms of population coupling, a concept recently proposed to understand the multi-neuron firing patterns of the visual cortex of mouse and monkey [21]. The population coupling characterizes the relationship of the activity of a single neuron with the population activity; this is because, the firing of one neuron is usually correlated with the firing pattern of other neurons. I further generalize the original population coupling to its higher order form, i.e., the relationship of pairwise firing with the population activity. I then derive the practical dimensionality reduction method for both types of population couplings, and test the method on different types of neural data, including ganglion cells in the salamander retina onto which a repeated natural movie was projected [17], and layer 2/3 as well as layer 5 cortical cells in the medial prefrontal cortex (MPC) of behaving rats [22]. In this paper, I develop a theoretical model of population coupling and its advanced form, to explain higher order correlations in the neural data. Methodologically, I propose the fast mean field method not only to learn the population couplings but also to evaluate the high order correlations. Note that this is computationally hard in a traditional maximum entropy model by using sampling-based method. Conceptually, I generalize the normal population coupling by introducing the second-order population coupling, which reveals interesting features from the data. First, it can explain a significant amount of three-cell correlations, and it works much better in cortical data than in retinal data. Second, the second-order population coupling matrix has distinct features in retinal and cortical data. The cortical one shows clear stripe-like structure while the retinal one has no such apparent structure. Altogether, this work marks a major step to understand the low-order representation of complex neural activity in both concepts and methods. 2 II. MODEL DESCRIPTION AND MEAN-FIELD METHODS For a neuronal population of size N , the neural spike trains of duration T are binned with temporal resolution τ , yielding M = (cid:100)T /τ(cid:101) samples of N -dimensional binary neural firing patterns. I use si = +1 to denote firing state of neuron i, and si = −1 for silent state. Neural responses to repeated stimulus (or the same behavioral tasks) vary substantially (so-called trial-to-trial variability) [23, 24]. To model the firing pattern statistics, I assign each firing pattern s a cost function (energy in statistical physics jargon) E(s), then the probability of observing one pattern s can be written as P (s) ∝ exp(−E(s)), where E(s) = −(cid:88) hisi −(cid:88) i i (cid:33) (cid:32)(cid:88) j(cid:54)=i Jisi sj . (1) This is the first low dimensional representation to be studied. High energy state s corresponds to low probability of observation. hi is the firing bias constraining the firing rate of neuron i, while Ji characterizes how strongly neuron i's spiking activity correlates with the population activity measured by the sum of other neurons' activity. I name Ji the first order population coupling (PopC1). Thus, only 2N parameters needs to be estimated from the neural data. This number of model parameters is much less than that in conventional maximum entropy model [17]. To model the high order correlation (e.g., three neuron firing correlation), I further generalize PopC1 to its advanced form, i.e., the second order population coupling, namely PopC2, describing the relationship of pairwise firing with the population activity, and the corresponding energy is given by E(s) = −(cid:88) hisi −(cid:88) i i<j (cid:33) (cid:32)(cid:88) k(cid:54)=i,j wijsisj sk , (2) where wij characterizes how strongly the firing state of the neuron pair (ij) correlates with the firing activities of other neurons. This term is expected to increase the prediction ability for modeling high order correlations in the neural data. Under the framework of PopC2, the total number of parameters to be estimated from the data is N + N (N − 1)/2. PopC1 and PopC2 have a clear neurobiological interpretation (for PopC1, see a recent study [21], and the results obtained under the PopC2 can also be experimentally tested), and moreover they can be interpreted in terms of functional interactions among neurons (as shown later). To find the model parameters as a low dimensional representation, I apply the maximum likelihood learning principle corresponding to maximizing the log-likelihood ln P (s) with respect to the parameters. The learning equation for PopC1 is given by (cid:32)(cid:68) (cid:69) (cid:32)(cid:88) (cid:68) si j(cid:54)=i data sisj −(cid:68) (cid:69) si (cid:33) (cid:69) (cid:68) −(cid:88) model , sisj data j(cid:54)=i ht+1 i = ht i + η J t+1 i = J t i + η (cid:33) , (cid:69) model (3a) (3b) where t and η denote the learning step and learning rate, respectively. The maximum likelihood learning shown here has a simple interpretation of minimizing the Kullback-Leibler divergence between the observation probability and the model probability [25, 26]. In an analogous way, one gets the learning equation for PopC2, (cid:32)(cid:88) (cid:68) k(cid:54)=i,j (cid:69) (cid:68) − (cid:88) k(cid:54)=i,j (cid:33) (cid:69) wt+1 ij = wt ij + η sisjsk data sisjsk model . (4) In the learning equations Eq. (3) and Eq. (4), the data dependent terms can be easily computed from the binned neural data. However, the model expectations of the firing rate (magnetization in statistical physics) and correlations are quite hard to evaluate without any approximations. Here I use the mean field method to tackle this difficulty. First, I write the energy term into a unified form, E(s) = −(cid:88) hisi −(cid:88) i a (cid:89) i∈∂a Γa si, (5) where a denotes the interaction index and ∂a denotes the neuron set involved in the interaction a. a = (ij) for PopC1 and (ijk) for PopC2. Therefore, PopC1 introduces the pairwise interaction as Γij = Ji + Jj, while PopC2 introduces the triplet interaction as Γijk = wij + wjk + wik. The multi-neuron interaction in the conventional Ising model is decomposed into first order or second order population coupling terms. This decomposition still maintains the functional heterogeneity of single neurons or neuron pairs, but reduces drastically the dimensionality of the neural representation for explaining high order correlations. In principle, one can combine PopC1 and PopC2 to predict both pairwise and triplet correlations. However, in this work, I focus on the pure effect of each type of population coupling. In fact, the conventional Ising model [27] can be recovered by setting Γij = Jij, which is pairwise interaction. The learning equation is derived similarly, and is run by reducing the deviation between the model pairwise correlation and the clamped one (computed from the data) [28]. Second, the statistical properties of the model (Eq. 5) can be analyzed by the cavity method in the mean field theory [29]. The self-consistent equations are written in the form of message passing (detailed derivations were given in Refs [30, 31], see also Appendix A) as hi + (cid:89) (cid:88) b∈∂i\a j∈∂b\i mi→a = tanh tanh −1 mb→i mb→i = tanh Γb mj→b,  , 3 (6a) (6b) (7) where ∂b\i denotes the member of interaction b except i, and ∂i\a denotes the interaction set i is involved in with a removed. mi→a is interpreted as the message passing from the neuron i to the interaction a it participates in, while mb→i is interpreted as the message passing from the interaction b to its member i. This iteration equation is also called the belief propagation (BP), which serves as the message passing algorithm for the statistical inference of the model parameters. Iteration of the message passing equation on the inferred model would converge to a fixed point corresponding to a global minimum of the free energy (in the cavity method approximation [31]) F ≡ − ln Z = −(cid:88) (cid:88) (cid:1). I define the function Hi(x) ≡ exhi(cid:81) ln Zi + a i (∂a − 1) ln Za, ron is − ln Zi = − ln(cid:80) ln(cid:0)1 + tanh Γa (cid:81) where Z is the normalization constant of the model probability P (s). The free energy contribution of one neu- x=±1 Hi(x) and the free energy contribution of one interaction is − ln Za = − ln cosh Γa − b∈∂i cosh Γb(1 + x mb→i). At the same time, the i∈∂a mi→a model firing rate and multi-neuron correlation can be estimated as mi = tanh tanh −1 mb→i (cid:32) (cid:88) tanh Γa +(cid:81) (cid:81) hi + b∈∂i (cid:33) Magnetization and correlation are defined as mi = (cid:104)si(cid:105) and Ca =(cid:10)(cid:81) 1 + tanh Γa Ca = i∈∂a mi→a i∈∂a mi→a . i∈∂a si , (cid:11), respectively. (8a) (8b) A brief derivation of Eq. (8) is given in Appendix A. Here the multi-neuron correlation is calculated directly from the cavity method approximation and expected to be accurate enough for current neural data analysis [28]. This is because, correlations under the model are evaluated taking into account nearest-neighbor interactions, rather than naive full independence among neurons. This approximation is expected to work well in a weakly-correlated neural population [5, 8], where long-range strong correlations do not develop. A similar application of this principle revealed a non-trivial geometrical structure of population codes in salamander retina [28]. Another advantage is the low computation cost. Both the free energy and the pairwise correlations can be estimated by the time complexity of the order O(N 2) for PopC1, and O(N 3) for triplet correlations in PopC2, which is one order of magnitude lower than the as S = −(cid:80) tractable model of PopC1 recently proposed in Ref. [32]. A more accurate expression could be derived from linear response theory [33] with much more expensive computational cost (increased by an order of magnitude (O(N )). To estimate the information carried by a neural population, one needs to compute the entropy, which is defined s P (s) ln P (s), and it measures the capacity of the neural population for information transmission. The more obvious variability the neural responses have, the larger the entropy value is. The entropy of the model can be estimated from the fixed point of the message passing equation. Based on the standard thermodynamic relation, S = −F + E, where E is the energy of the neural population and given by 4 (9a) (9b) (9c) (9d)  mj→b (cid:89) j∈∂b\i a i hi ∆Ei + (∂a − 1)∆Ea, x=±1 Gi(x) E = −(cid:88) (cid:88) (cid:80) x=±1 xHi(x) +(cid:80) (cid:80) tanh Γa +(cid:81) x=±1 Hi(x) (cid:81) Γb sinh Γb(1 + x mb→i) + xΓb cosh Γb(1 − tanh2 Γb) i∈∂a mi→a i∈∂a mi→a 1 + tanh Γa , , Gi(x) = exhi cosh Γa(1 + x ma→i). ∆Ei = ∆Ea = Γa (cid:88) × (cid:89) b∈∂i a∈∂i\b The basic procedure to infer population couplings is given as follows. At the beginning, all model parameters are assigned zero value. It is followed by three steps: (i) Messages are initialized randomly and uniformly in the interval (−1, 1). (ii) Eq. (6) are then run until converged, and the magnetizations as well as multi-neuron correlations are estimated using Eq. (8). (iii) The estimated magnetizations and correlations are used at each gradient ascent learning step (Eq. (3) or Eq. (4)). When one gradient learning step is finished, another step starts by repeating the above procedure (from (i) to (iii)). To learn the higher order population coupling, the damping technique is used to avoid oscillation behavior, i.e., wnew = γwnew + (1 − γ)wold where γ is the damping factor taking a small value. The inferred model can also be used to generate the distribution of spike synchrony, i.e., the probability of K simultaneous spikes. This distribution can be estimated by using Monte Carlo (MC) simulation on the model. The standard procedure goes as follows. The simulation starts from a random initial configuration s0, and tries to search for the low energy state, then the energy is lowered by a series of elementary updates, and for each elementary update, N proposed neuronal state flips are carried out. That is, the transition probability from state s to s(cid:48) with only si flipped (s(cid:48) j∈∂a\i sj. The equilibrium samples are collected after sufficient thermal equilibration. These samples (a total of 20000 samples in simulations) are finally used to estimate the distribution of spike synchrony. i = −si) is expressed as e−2siHi where Hi = hi +(cid:80) a∈∂i Γa (cid:81) III. RESULTS By using the mean field method, I first test both types of population couplings on the retina data, which is the spike train of 160 ganglion cells in a small patch of the salamander retina [17]. The retina was stimulated with a repeated natural movie. The spike train data is binned with the bin size equal to 20ms reflecting the temporal correlation time scale, yielding about 282744 binary firing patterns for data modeling. I then test the same concepts on the cortical data of behaving rats. Rats performed the odor-place matching working memory during one task session, and spiking activities of 117 cells in both superficial layer 2/3 and deep layer (layer 5) of medial prefrontal cortex were simultaneously recorded (for detailed experiments, see Ref. [22]). One task session consists of about 40 trials, yielding a spike train of these cortical cells binned with the temporal resolution τ = 20ms (a total of 140596 firing patterns). A. Inference performances on the retinal data Fig. 1 reports the inference result on a network example of 40 neurons selected randomly from the original dataset. The firing rate is predicted faithfully by the model using either MC or BP (Fig. 1 (A)). Inferring only PopC1, one could predict about 74.10% of entire pairwise correlation (a precision criterion is set to 7.0 × 10−3 in this paper) (Fig. 1 (B)). This means that 74.10% of the whole correlation set have the absolute value of the deviation between 5 FIG. 1: (Color online) Inference performances of PopC1 on the retinal data (N = 40, one typical example). (A) Firing rate in the data is reproduced by the model. (B) Two-cell correlations are explained partially by PopC1 (74.10%). A Monte-Carlo (MC) sampling of the model yields similar results to belief propagation (BP), which is much faster. (C) From the MC samples, three-cell correlations can also be estimated. (D) Probability of K synchronous spiking under the model is compared with that of the data. a ) smaller than the precision criterion. Using the sampled the predicted correlation and measured one (C pred configurations of neural firing activity from the MC simulation, one could also predict three-cell correlations (Fig. 1 (C)), whereas, the prediction fraction can be improved by a significant amount after introducing PopC2, as I shall show later. In addition, fitting only 2N model parameters in PopC1 analysis could not predict the tail of spike synchrony distribution (Fig. 1 (D)); this is expected as no higher order interaction terms are included in the model, and rare events of large K spikes are also difficult to observe in a finite sampling during MC simulations. a − C ms The inference results of PopC2 are given in Fig. 2. Note that, by considering the correlation between the pairwise firing activity and the global population activity, i.e., the second order population coupling, the three-cell correlation could be predicted partially (64.44%), and this fraction is much larger than that of PopC1 (Fig. 1 (C)). This is due to the specific structure of PopC2, which incorporates explicitly three-cell correlations into the construction of couplings (Eq. (4)). Technically, the mean-field theory for PopC2 avoids the slow sampling and evaluates the high order correlations in a fast way. Alternatively, one could fit the data using the conventional Ising model [27] with the same number of model parameters as PopC2, whereas, the three-cell correlations are hard to predict using MC samplings, and a similar phenomenon was also observed in a previous work for modeling pairwise correlations [33]. Therefore I speculate that PopC2 acts as a key circuit variable for third order correlations. The interaction matrix of {wij} reveals how important each pair of neurons is for the entire population activity (emergent functional state of the whole network). As shown in Fig. 2 (C), PopC2 matrix has no apparent structure of organization, i.e., each neuron can be paired with both positive and negative couplings. Some pairs have large negative PopC2, suggesting that these components are anti-correlated with the population activity. That is to say, the activity of these neuron-pairs is not synchronized to the population activity characterized by the summed activity over all neurons except these pairs. In the network, there also exist positive PopC2s, which shows that these neuron-pairs are positively correlated with the population in neural activity. The interaction matrix shown here may be related to two-cell correlation(model)0.70.750.80.850.90.951two-cell correlation (data)0.70.750.80.850.90.951BP 74.10%MCequality(B)P(K)10−510−410−310−210−1100K024681012141618DataMC(D) mean firing rate (model)−0.95−0.9−0.85−0.8mean firing rate (data)−1−0.95−0.9−0.85−0.8BPMCequality(A)three-cell correlation(model)−1−0.9−0.8−0.7−0.6three-cell correlation (data)−1−0.9−0.8−0.7−0.6MC 46.31% equality(C) 6 FIG. 2: (Color online) Inference performances of PopC2 on the retinal data (N = 40, one typical example). (A) Firing rate in the data is reproduced by the model. (B) Three-cell correlations are explained partially by PopC2 (64.44%). (C) Interaction matrix for PopC2. (D) Probability of K synchronous spiking under the model is compared with that of the data. FIG. 3: (Color online) Inference performances of Ising model on the retinal data (N = 40, one typical example) compared with population coupling. (a) Two-neuron interaction matrix. (b) Probability of K synchronous spiking under the model is compared with that of the data. the revealed overlapping modular structure of retinal neuron interactions [8, 14]. In this structure, neurons interact locally with their adjacent neurons, and in particular this feature is scalable and applicable for larger networks. It seems that one individual neuron does not impact directly the entire population, and a small group of neighboring neurons have similar visual feature selectivity [34]. This result is also consistent with two-neuron interaction map of the conventional Ising model (Fig. 3 (a)). Note that in functional interpretation, these two-neuron interactions are inherently different from PopC2, which is designed to explain high-order correlations by using less model parameters than necessary. three-cell correlation (model)−1−0.95−0.9−0.85−0.8−0.75−0.7−0.65three-cell correlation (data)−1−0.95−0.9−0.85−0.8−0.75−0.7−0.652F1BP 64.44%equality(B) mean firing rate (model)−1−0.95−0.9−0.85 mean firing rate (data)−1−0.95−0.9−0.852F1F2(A) P(K)10−510−410−310−210−1100K02468101214DataMC(D)(C) neuron index510152025303540 −0.01−0.008−0.006−0.004−0.00200.002neuron index510152025303540510152025303540(a) neuron index0510152025303540 −0.100.10.20.30.4neuron index05101520253035400510152025303540 (b)P(K)10−510−410−310−210−1100 K024681012 dataPopC1PopC2 Ising 7 FIG. 4: (Color online) Multi-information and prediction fraction of correlations under the simplified model for the retinal network. The result is averaged over 10 network samples for each N . (a) Multi-information (in bits) versus the network size N . (b) The prediction fraction ρ versus the network size. PopC1 is used to predict two-cell correlations, and PopC2 is used to predict three-cell correlations. PopC2 behaves better than PopC1 in predicting the spike synchrony distribution (Fig. 2 (D)) in the small K regime (the prediction is improved from K = 4 for PopC1 to K = 8 for PopC2). An intuitive explanation is that PopC2 introduces equivalently triplet interactions among neurons, and it is known that high order interactions are necessary for generating widespread population activity [12]. However, PopC2 overestimates the distribution when rare events of synchronous spiking are considered. This may be related to the difficulty of obtaining sufficient equilibrium samples of the model, especially those samples with large population activity. The spike synchrony distribution is also compared with that obtained under Ising model (Fig. 3 (b)). Different performances are related to the multi-information measure of neural population explained below. Sind−Smodel, in which Sind =(cid:80) The amount of statistical structure in the neural data due to introducing interactions among neurons can be (cid:80) measured by the multi-information [5]. I first introduce an independent model where only the firing rates of individual neurons are fitted and the corresponding entropy is defined as Sind. The multi-information is then defined as I(N ) = x=±1 S((1+mix)/2), where S(u) = −u ln u, and Smodel is assumed to be an upper bound to the true entropy. The true entropy for large populations is difficult to estimate since it requires including all possible interactions among neurons. However, the model entropy with low order interaction parameters could be an approximate information capacity for the neural population, which depends on how significant the higher order correlations are in the population. i Fig. 4 (a) shows the multi-information as a function of the network size. PopC1 and PopC2 are compared with the Ising model [33], which reconstructs faithfully the pairwise correlations. PopC2 improves significantly over PopC1 in capturing the information content of the network, but its multi-information is still below that of the Ising model, which is much more evident for larger network size. This is expected, because only part of third order correlations are captured by PopC2, while the Ising model describes accurately the entire pairwise correlation profile which may be the main contributor to the collective behavior observed in the population. However, PopC2 provides us an easy way to understand the higher order correlation, while in the Ising model, it is computationally difficult to estimate the higher order correlations. The average prediction fraction of correlations by PopC1 and PopC2 is plotted in Fig. 4 (b). PopC1 predicts more than 75% of the pairwise correlations, while PopC2 predicts more than 54% of the triplet correlations. The prediction fraction changes slightly with the network size. B. Inference performances on the cortical data To show the inference performance of both types of population couplings on the cortical data, I randomly select a typical network example of 40 neurons from the original dataset, and then apply the computation scheme to this typical example. Results are shown in Fig. 5. Surprisingly, the simplified PopC1 is able to capture as high as 99.23% of pairwise correlations, implying that when a rat performed working memory tasks, there exists a simplified model to describe emergent functional states in the medial prefrontal cortical circuit. Moreover, MC sampling of the PopC1 model also predicts well the spike synchrony distribution (Fig. 5 (D)). This is very different from that observed in the Multi-Information00.20.40.60.811.21.41.6 N30405060708090IsingPopC1PopC2(a) ρ0.50.550.60.650.70.750.8 N30405060708090PopC1PopC2(b) 8 FIG. 5: (Color online) Inference performances of PopC1 on the cortical data (N = 40, one typical example). (A) Firing rate in the data is reproduced by the model. (B) Two-cell correlations are explained partially by PopC1 (99.23%). A Monte-Carlo (MC) sampling of the model yields similar results to belief propagation (BP). (C) From the MC samples, three-cell correlations can also be estimated. (D) Probability of K synchronous spiking under the model is compared with that of the data. retinal data. In this sense, the MPC circuit is simple in its functional states when the subject is performing specified tasks. More interesting circuit features are revealed by PopC2, which is shown in Fig. 6. About 94.79% of three-cell correlations are explained by PopC2 in the MPC circuit. The interaction matrix of PopC2 in Fig. 6 (C) shows a clear non-local structure in the cortical circuit (stripe-like structure). That is, some neurons interact strongly with nearly all the other neurons in the selected population, and these interactions have nearly identical strength of PopC2. Such neurons having stripe-like structure in the PopC2 matrix may receive a large number of excitatory inputs from pyramidal neurons [22], and thus play a key role in shaping the collective spiking behavior during the working memory task. The non-local effects are consistent with findings reported in the original experimental paper (cross-correlogram analysis) [22] and the two-neuron interaction map under Ising model (Fig. 7 (a)). Thus, to some extent, PopC2 may reflect intrinsic connectivity in the cortical circuit, although the relationship between functional connections and anatomical connections has not yet been well established [35]. Lastly, PopC2 overestimates the tail of the spike synchrony distribution (Fig. 6 (D)), which may be caused by the sampling difficulty of the inferred model (a model with triplet interactions among its elements). The spike synchrony distribution of Ising model is also compared (Fig. 7 (b)). Multi-information versus the cortical network size is plotted in Fig. 8 (a). In the cortical circuit, PopC2 behaves comparably with the Ising model; even for some network size (N = 50), it reports a higher information content than the Ising model in the randomly selected subpopulations, which may be caused by the nature of the selected neurons (e.g., inhibitory interneurons [22], and they have stripe-like structure in the PopC2 matrix). Note that PopC1 gives an information close to zero for small network sizes, suggesting that by introducing PopC1, one could not increase significantly the amount of statistical structure in the network activity explained by the model. However, the multi- information of PopC1 grows with the network size, indicating that the role of PopC1 would be significant for larger neural populations. Fig. 8 (b) reports the prediction fraction of the correlation profile by applying PopC1 and PopC2. Both population couplings can capture over 90% of correlations, which is significantly different from that observed in the retinal data. mean firing rate (model)−0.9−0.8−0.7−0.6−0.5−0.4mean firing rate (data)−1−0.9−0.8−0.7−0.6−0.5−0.423F1BPMCequality(A) two-cell correlation(model)0.20.40.60.81two-cell correlation (data)0.20.40.60.81BP 99.23%MCequality(B)three-cell correlation(model)−1−0.8−0.6−0.4−0.2three-cell correlation (data)−1−0.8−0.6−0.4−0.2MC 89.51% equality(C)P(K)10−510−410−310−210−1100K0123456789DataMC(D) 9 FIG. 6: (Color online) Inference performances of PopC2 on the cortical data (N = 40, one typical example). (A) Firing rate in the data is reproduced by the model. (B) Three-cell correlations are explained partially by PopC2 (94.79%). (C) Interaction matrix for PopC2. (D) Probability of K synchronous spiking under the model is compared with that of the data. FIG. 7: (Color online) Inference performances of Ising model on the cortical data (N = 40, one typical example) compared with population coupling. (a) Two-neuron interaction matrix. (b) Probability of K synchronous spiking under the model is compared with that of the data. IV. DISCUSSION The emergent properties of the neural code arise from interactions among individual neurons. A complete charac- terization of the population activity is difficult, because on the one hand, the number of potential interactions suffers from a combinatorial explosion, on the other hand, the collective behavior at the network level would become much more complex as the network size grows. In this paper, I develop a theoretical framework to understand how pairwise or higher order correlations arise and the basic circuit variables corresponding to these correlation structures. The three-cell correlation (model)−1−0.9−0.8−0.7−0.6−0.5−0.4−0.3−0.2three-cell correlation (data)−1−0.9−0.8−0.7−0.6−0.5−0.4−0.3−0.22F1BP 94.79%equality(B) P(K)10−510−410−310−210−1100K02468DataMC(D) mean firing rate (model)−1−0.9−0.8−0.7−0.6−0.5−0.4 mean firing rate (data)−1−0.9−0.8−0.7−0.6−0.5−0.42F1F2(A)(C) neuron index510152025303540 −0.0015−0.001−0.000500.0005neuron index510152025303540510152025303540(b)P(K)10−510−410−310−210−1100 K02468 dataPopC1PopC2 Ising(a)neuron index0510152025303540 00.020.040.060.08neuron index05101520253035400510152025303540 10 FIG. 8: (Color online) Multi-information and prediction fraction of correlations under the simplified model for the cortical network. The result is averaged over 10 network samples for each N . (a) Multi-information (in bits) versus the network size N . (b) The prediction fraction ρ versus the network size. PopC1 is used to predict two-cell correlations, and PopC2 is used to predict three-cell correlations. model is based on the concept of population coupling, characterizing the relationship between local firing activity of individual neuron or neuron-pair and the global neural activity. An advantage is that, it provides a low dimensional and neurobiologically interpretable representation to understand the functional interaction between neurons and their correlation structures. In particular, the concept of population coupling and the associated mean field method used in this paper offer an easy way to evaluate higher order correlations, while the usual sampling method is computationally hard and traditional models (e.g., Ising model) lack a direct interpretation of higher order correlations in terms of simplified (population) couplings. With the mean field method, the concept of population coupling is tested on two different types of neural data. One is the firing neural activities of retinal ganglion cells under natural movie stimuli. The other is the population activities of medial prefrontal cortex when a rat was performing odor-place matching working memory tasks. For the retinal data, on average PopC1 accounts for more than 75% of pairwise correlations, and PopC2 accounts for over 54% of three-cell correlations. The interaction matrix of PopC2 contains information about the functional interaction features in the retinal circuitry. It seems that a retinal neuron can be paired with not only negatively strong couplings, but also slightly positive couplings. Only a few pairs of neurons have strong correlations with the global activity of the population. To describe the spike synchrony distribution, PopC2 performs better than PopC1, nevertheless, both of them could not capture the trend of the tail (rare events related to higher order interactions existing in the network). This is not surprising, because PopC1 and PopC2 are simplified descriptions of the orig- inal high dimensional neural activity, taking the trade-off between the computation complexity and the description goodness. To extract the statistical structure embedded in the neural population, PopC2 improves significantly over PopC1, and has further additional benefit of describing the third-order correlations observed in the data, as PopC2 could be used to construct triplet interactions among neurons, although direct constructing all possible triplet interactions is extremely computationally difficult. Unlike the retinal circuit, the cortical circuit yields a much smaller absolute value of the multi-information, implying that no significant higher order correlations (interactions) were present in the neural circuit when the circuit was car- rying out task-related information processing rather than encoding well-structured stimuli (as in the retinal network). This also explains why a simplified description such as PopC1 and PopC2 is accurate enough to capture the main features of the population activity, including the spike synchrony distribution. The inferred model on the cortical data reveals a different interaction map from that of the retinal circuit. In the cortical circuit, neurons form the stripe-like structure in the interaction matrix, suggesting that these neurons may receive a large number of excitatory inputs [22]. These inputs may come from different layers of cortex, and they can execute top-down or bottom-up information processing, thus modulate the global brain state in the target cortex during behavioral tasks. Before summary of this work, I made some discussions about two relevant recent studies on population coupling (see notes added). Ref. [32] modeled directly the joint probability distribution of individual neural response and population rate (the number of neurons simultaneously active) by linear coupling and complete coupling models. The linear coupling reproduces separately the distribution of individual neuronal state and the population rate distribution, and their couplings, while PopC1 introduced in my work reproduces mean firing rate and the correlation between Multi-information−0.0200.020.040.060.080.1 N30405060708090 IsingPopC1PopC2(a) ρ0.880.90.920.940.960.98 N30405060708090PopC1PopC2(b) 11 individual neuronal state and the background population activity (except the neuron itself). Note that PopC1 does not model population rate distribution explicitly, which is hard to interpret in terms of functional connectivity. The complete-coupling model reproduces the joint probability distributions between the response of each neuron and the population rate, from which it is hard to conclude that the high-order interactions responsible for high-order correlations can be interpreted and tested. However, PopC2 reproduces mean firing rate and the correlation between neuron-pair activities and the background population activity (except the neuron-pair itself), and thus explains high- order correlations by an energy model. Furthermore, in this sense, this work overcame a weakness pointed out in another independent later work of population coupling [36], which fitted directly the population rate distribution and the firing probability for each neuron conditioned on the population rate, and analogously the corresponding model parameters can not be readily interpretable in a biological setting. Due to intrinsic difference in model definitions, these two relevant works have nice properties of studying tuning curves of individual neurons to the population rate, and sampling from the model to reproduce the population synchrony distribution. In summary, I develop a theoretical model of population coupling and its advanced form, to relate the correlation profile in the neural collective activity to the basic circuit variables. The practical dimensional reduction method is tested on different types of neural data, and specific features of neural circuit are revealed. This model aiming at describing high order correlations with a low order representation, is expected to be useful for modeling big neural data. Note that the interaction matrices shown in Fig. 2 and Fig. 6 are qualitatively robust to changes of the data size to only the first half (data not shown), verifying that the revealed features are not an artifact of overfitting. However, it still deserves further studies by introducing regularization in the learning equation. It is also very interesting to incorporate more physiologically plausible parameters to explain how the collective spiking behavior arises from the microscopic interactions among the basic units. Another interesting study is to clarify the role of higher order correlations in decoding performances based on maximum likelihood principles [37, 38]. Acknowledgments I am grateful to Shigeyoshi Fujisawa and Michael J Berry for sharing me the cortical and retinal data, respectively. I also thank Hideaki Shimazaki and Taro Toyoizumi for stimulating discussions. This work was supported by the program for Brain Mapping by Integrated Neurotechnologies for Disease Studies (Brain/MINDS) from Japan Agency for Medical Research and development, AMED. note added. -- After I submitted this work to arXiv:1602.08299, I became aware of Ref. [32] (arXiv:1606.08889), and later Ref. [36] (bioRxiv, 2016). Discussions about these two recent relevant works are made in the last section of this paper. Appendix A: Derivation of mean-field equations In this appendix, (cid:81) I give a simple derivation of mean-field equations given in Sec. II. More details can First, after removing an interaction a, one defines the cavity probability be obtained from Refs [30, 31]. Pi→a(si) = ehisi Pb→i(si), where the product comes from the physical meaning of the second kind of Zi j∈∂b sj(cid:81) cavity probability, namely Pb→i(si) defined as the cavity probability when only the connection from b to i is (cid:81) Pb→i(si) is thus formulated as retained while other neighbors of i are removed (so-called cavity probability). it follows that the cavity magnetization j∈∂b\i Pj→b(sj). With these two probabilities, b∈∂i\a (cid:80) (cid:16) mi→a =(cid:80) mb→i by mb→i = tanh(ub→i) =(cid:80) siPi→a(si) = tanh sj :j∈∂b\i eΓb si hi +(cid:80) si Pb→i(si). si b∈∂i\a ub→i , where ub→i is named cavity bias in physics [29]. It is related to More specifically, the cavity magnetization is derived as follows, (cid:17) ehi(cid:81) ehi(cid:81)  , mi→a = Pi→a(+1) − Pi→a(−1) = −hi−2(cid:80) −hi−2(cid:80) hi + (cid:88) ehi − e ehi + e = = tanh b∈∂i\a ub→i b∈∂i\a ub→i ub→i b∈∂i\a Pb→i(+1) − e−hi(cid:81) Pb→i(+1) + e−hi(cid:81) b∈∂i\a b∈∂i\a b∈∂i\a b∈∂i\a Pb→i(−1) Pb→i(−1) (A1a) where I have used the definition ub→i ≡ 1 interaction, 2 ln Pb→i(+1) Pb→i(−1) . Next, I show how to derive the cavity bias. First, for pairwise 12 ub→i = = 1 2 1 2 ln ln = tanh eΓb Pj→b(+1) + e−Γb Pj→b(−1) e−Γb Pj→b(+1) + eΓb Pj→b(−1) 1 + tanh Γbmj→b 1 − tanh Γbmj→b −1 (tanh Γbmj→b) , (A2a) where I have used the parameterization Pj→b(sj) = 1+sj mj→b for triplet interaction, 2 , and the mathematical identity e2z = 1+tanh z 1−tanh z . Similarly, ub→i = = 1 2 1 2 ln ln = tanh eΓb [Pj→b(+1)Pk→b(+1) + Pj→b(−1)Pk→b(−1)] + e−Γb [Pj→b(+1)Pk→b(−1) + Pj→b(−1)Pk→b(+1)] e−Γb [Pj→b(+1)Pk→b(+1) + Pj→b(−1)Pk→b(−1)] + eΓb [Pj→b(+1)Pk→b(−1) + Pj→b(−1)Pk→b(+1)] 1 + e−2Γb 1−mj→bmk→b e−2Γb + 1−mj→bmk→b −1 (tanh Γbmj→bmk→b) . 1+mj→bmk→b 1+mj→bmk→b These results are written in a compact form as Eq. (6) in the main text. Similarly, the single neuron magnetization mi is obtained via mi = 1 Zi si:i∈∂a Pb→i(si), where Zi is a normalization constant, and the multi-neuron correlation Ca = 1 i∈∂a Pi→a(si), where Za Za is a normalization constant. Note that Zi and Za are also related to the free energy contribution of single neuron and neuronal interaction [30], respectively. The full (non-cavity) magnetization can be derived in a similar manner to (cid:1). In detail, the two-point correlation Ca (∂a = 2) is computed as follows, Eq. (A1), as mi = tanh(cid:0)hi +(cid:80) si i∈∂a sieΓa b∈∂i ub→i (cid:80) (cid:81) siehisi(cid:81) i∈∂a si(cid:81) (cid:81) b∈∂i (cid:80) (cid:80) (cid:80) si,sj Ca ≡ (cid:104)sisj(cid:105) = sisjeΓasisj Pi→a(si)Pj→a(sj) eΓasisj Pi→a(si)Pj→a(sj) si,sj 1+mj→ami→a e2Γa − 1−mj→ami→a e2Γa + 1−mj→ami→a tanh Γa + mi→amj→a 1 + tanh Γami→amj→a 1+mj→ami→a . = = (cid:80) (cid:80) si,sj ,sk si,sj ,sk 1+mj→ami→amk→a e2Γa − 1−mj→ami→amk→a e2Γa + 1−mj→ami→amk→a tanh Γa + mi→amj→amk→a 1 + tanh Γami→amj→amk→a 1+mj→ami→amk→a . = = (A3a) (A4a) (A5a) Analogously, three-point correlation (∂a = 3) can be evaluated as Ca ≡ (cid:104)sisjsk(cid:105) = sisjskeΓasisj sk Pi→a(si)Pj→a(sj)Pk→a(sk) eΓasisj sk Pi→a(si)Pj→a(sj)Pk→a(sk) These results are written in a compact form as Eq. (8b) in the main text. Finally, the partition function Zi for pairwise interaction is computed as follows, Zi = ehi (cid:89) = ehi (cid:89) b∈∂i Pb→i(+1) + e−hi (cid:89) cosh Γb(1 + tanh Γbmj→b) + e−hi (cid:89) Pb→i(−1) b∈∂i b∈∂i b∈∂i cosh Γb(1 − tanh Γbmj→b), (A6a) where I have used Pb→i(si) = eΓbsiPj→b(+1) + e−ΓbsiPj→b(−1) and the magnetization parameterization of Pj→b. For triplet interaction, Pb→i(si) = eΓbsi[Pj→b(+1)Pk→b(+1) + Pj→b(−1)Pk→b(−1)] + e−Γbsi [Pj→b(+1)Pk→b(−1) + Pj→b(−1)Pk→b(+1)], and the corresponding Zi = ehi(cid:81) b∈∂i cosh Γb(1 + tanh Γbmj→bmk→b) + e−hi(cid:81) 13 b∈∂i cosh Γb(1 − tanh Γbmj→bmk→b). Following the same line, the partition function Za is evaluated for pairwise interaction as Za = eΓasisj Pi→a(si)Pj→a(sj) = cosh Γa[1 + tanh Γami→amj→a]. (cid:88) si,sj (cid:88) (A7a) (A8a) Similarly, the partition function Za for triplet interaction can be computed as Za = eΓasisj sk Pi→a(si)Pj→a(sj)Pk→a(sk) which is exactly the compact equation given in the main text. si,sj ,sk = cosh Γa[1 + tanh Γami→amj→amk→a], [1] R. Q. Quiroga and S. Panzeri. Extracting information from neuronal populations: information theory and decoding approaches. Nat Rev Neurosci, 10:173, 2009. [2] R. Yuste. From the neuron doctrine to neural networks. Nat Rev Neurosci, 16:487, 2015. [3] I. H. Stevenson and K. P. Kording. How advances in neural recording affect data analysis. Nat Rev Neurosci, 14:139, 2011. [4] M. R. Cohen and A. Kohn. Measuring and interpreting neuronal correlations. Nat Neurosci, 14:811, 2011. [5] E. Schneidman, M. J. Berry, R. Segev, and W. Bialek. Weak pairwise correlations imply strongly correlated network states in a neural population. Nature, 440:1007, 2006. [6] S. Cocco, S. Leibler, and R. Monasson. Neuronal couplings between retinal ganglion cells inferred by efficient inverse statistical physics methods. Proc. Natl. Acad. Sci. USA, 106:14058, 2009. [7] J. H. Macke, M. Opper, and M. Bethge. Common input explains higher-order correlations and entropy in a simple model of neural population activity. Phys. Rev. Lett, 106:208102, 2011. [8] Elad Ganmor, Ronen Segev, and Elad Schneidman. The architecture of functional interaction networks in the retina. The Journal of Neuroscience, 31:3044, 2011. [9] L. Martignon, G. Deco, K. Laskey, M. Diamond, W. Freiwald, and E. Vaadia. Neural coding: higher-order temporal patterns in the neurostatistics of cell assemblies. Neural Computation, 12:2621, 2000. [10] M. J. Schnitzer and M. Meister. Multineuronal firing patterns in the signal from eye to brain. Neuron, 37:499, 2003. [11] I. E. Ohiorhenuan, F. Mechler, K. P. Purpura, A. M. Schmid, Q. Hu, and J. D. Victor. Sparse coding and high-order correlations in fine-scale cortical networks. Nature, 466:617, 2010. [12] S.-I. Amari, H. Nakahara, S. Wu, and Y. Sakai. Synchronous firing and higher-order interactions in neuron pool. Neural Computation, 15:127, 2003. [13] F. Montani, E. Phoka, M. Portesi, and S. R. Schultz. Statistical modelling of higher-order correlations in pools of neural activity. Physica A: Statistical Mechanics and its Applications, 392:3066, 2013. [14] E. Ganmor, R. Segev, and E. Schneidman. Sparse low-order interaction network underlies a highly correlated and learnable neural population code. Proc. Natl. Acad. Sci. USA, 108:9679, 2011. [15] Shan Yu, Hongdian Yang, Hiroyuki Nakahara, Gustavo S. Santos, Danko Nikoli´c, and Dietmar Plenz. Higher-order interactions characterized in cortical activity. The Journal of Neuroscience, 31:17514, 2011. [16] G. Tkacik, O. Marre, T. Mora, D. Amodei, M. J. Berry II, and W. Bialek. The simplest maximum entropy model for collective behavior in a neural network. Journal of Statistical Mechanics: Theory and Experiment, 2013:P03011, 2013. [17] G. Tkacik, O. Marre, D. Amodei, E. Schneidman, W. Bialek, and M. J. Berry II. Searching for collective behavior in a large network of sensory neurons. PLoS Comput Biol, 10:e1003408, 2014. [18] U. Koster, J. Sohl-Dickstein, C. M. Gray, and B. A. Olshausen. Modeling higher-order correlations within cortical micro- columns. PLoS Comput Biol, 10:e1003684, 2014. [19] Haiping Huang. Effects of hidden nodes on network structure inference. Journal of Physics A: Mathematical and Theoretical, 48:355002, 2015. [20] H. Shimazaki, K. Sadeghi, T. Ishikawa, Y. Ikegaya, and T. Toyoizumi. Simultaneous silence organizes structured higher- order interactions in neural populations. Scientific Reports, 5:9821, 2015. [21] M. Okun, N. A. Steinmetz, L. Cossell, M. F. Iacaruso, H. Ko, P. Bartho, T. Moore, S. B. Hofer, T. D. Mrsic-Flogel, M. Carandini, and K. D. Harris. Diverse coupling of neurons to populations in sensory cortex. Nature, 521:511, 2015. [22] S. Fujisawa, A. Amarasingham, M. T. Harrison, and G. Buzsaki. Behavior-dependent short-term assembly dynamics in the medial prefrontal cortex. Nat Neurosci, 11:823, 2008. [23] Rob R. de Ruyter van Steveninck, Geoffrey D. Lewen, Steven P. Strong, Roland Koberle, and William Bialek. Repro- ducibility and variability in neural spike trains. Science, 275:1805, 1997. [24] Michael Okun, Pierre Yger, Stephan L Marguet, Florian Gerard-Mercier, Andrea Benucci, Steffen Katzner, Laura Busse, Matteo Carandini, and Kenneth D Harris. Population rate dynamics and multineuron firing patterns in sensory cortex. The Journal of Neuroscience, 32:17108, 2012. 14 [25] Simona Cocco and R´emi Monasson. Adaptive cluster expansion for the inverse ising problem: convergence, algorithm and tests. J. Stat. Phys, 147:252, 2012. [26] H. Huang. Sparse hopfield network reconstruction with (cid:96)1 regularization. Eur. Phys. J. B, 86:484, 2013. [27] Y. Roudi, E. Aurell, and J. Hertz. Statistical physics of pairwise probability models. Front. Comput. Neurosci, 3:1, 2009. [28] Haiping Huang and Taro Toyoizumi. Clustering of neural code words revealed by a first-order phase transition. Phys. Rev. E, 93:062416, 2016. [29] M. M´ezard and G. Parisi. The bethe lattice spin glass revisited. Eur. Phys. J. B, 20:217, 2001. [30] H. Huang and H. Zhou. Cavity approach to the sourlas code system. Phys. Rev. E, 80:056113, 2009. [31] M. M´ezard and A. Montanari. Information, Physics, and Computation. Oxford University Press, Oxford, 2009. [32] Christophe Gardella, Olivier Marre, and Thierry Mora. A Tractable Method for Describing Complex Couplings between Neurons and Population Rate. eneuro, 3:e0160 -- 15.2016, 2016. arXiv:1606.08889. [33] H. Huang and H. Zhou. Counting solutions from finite samplings. Phys. Rev. E, 85:026118, 2012. [34] Ko Ho, Hofer Sonja B., Pichler Bruno, Buchanan Katherine A., Sjostrom P. Jesper, and Mrsic-Flogel Thomas D. Functional specificity of local synaptic connections in neocortical networks. Nature, 473:87, 2011. [35] Dimitri Yatsenko, Josic Kresimir, Ecker Alexander S., Froudarakis Emmanouil, Cotton R. James, and Tolias Andreas S. Improved Estimation and Interpretation of Correlations in Neural Circuits. PLoS Comput Biol, 11:e1004083, 2015. [36] Cian O'Donnell, J Tiago Goncalves, Nick Whiteley, Carlos Portera-Cailliau, and Terrence J Sejnowski. The population tracking model: A simple, scalable statistical model for neural population data. Neural Computation, 29:50 -- 93, 2017. [37] Michael T. Schaub and Simon R. Schultz. The ising decoder: reading out the activity of large neural ensembles. J Comput Neurosci, 32:101, 2012. [38] Greg Schwartz, Jakob Macke, Dario Amodei, Hanlin Tang, and Michael J. Berry. Low error discrimination using a correlated population code. J Neurophysiol, 108:1069, 2012.
1707.05952
1
1707
2017-07-19T06:32:34
Criticality in the brain: A synthesis of neurobiology, models and cognition
[ "q-bio.NC" ]
Cognitive function requires the coordination of neural activity across many scales, from neurons and circuits to large-scale networks. As such, it is unlikely that an explanatory framework focused upon any single scale will yield a comprehensive theory of brain activity and cognitive function. Modelling and analysis methods for neuroscience should aim to accommodate multiscale phenomena. Emerging research now suggests that multi-scale processes in the brain arise from so-called critical phenomena that occur very broadly in the natural world. Criticality arises in complex systems perched between order and disorder, and is marked by fluctuations that do not have any privileged spatial or temporal scale. We review the core nature of criticality, the evidence supporting its role in neural systems and its explanatory potential in brain health and disease.
q-bio.NC
q-bio
Criticality in the brain Cocchi et al. 1 Criticality in the brain: A synthesis of neurobiology, models and cognition Luca Cocchi1*, Leonardo L. Gollo1, Andrew Zalesky2, Michael Breakspear1 1 Queensland Institute for Medical Research, Brisbane, Australia. 2 Melbourne Neuropsychiatry Centre, The University of Melbourne, Melbourne, Australia. Abstract Cognitive function requires the coordination of neural activity across many scales, from neurons and circuits to large-scale networks. As such, it is unlikely that an explanatory framework focused upon any single scale will yield a comprehensive theory of brain activity and cognitive function. Modelling and analysis methods for neuroscience should aim to accommodate multiscale phenomena. Emerging research now suggests that multi-scale processes in the brain arise from so-called critical phenomena that occur very broadly in the natural world. Criticality arises in complex systems perched between order and disorder, and is marked by fluctuations that do not have any privileged spatial or temporal scale. We review the core nature of criticality, the evidence supporting its role in neural systems and its explanatory potential in brain health and disease. Keywords: Bifurcations; metastability; multistability; dynamics; power-law; cognition. Highlights  Criticality is a wide-spread phenomenon in natural systems  Criticality provides a unifying framework to model and understand brain activity and cognitive function  Substantial evidence now supports the hypothesis that the brain operates near criticality  We review the role of criticality in healthy and pathological brain dynamics  Caveats and pitfalls regarding the assessment of criticality in the brain are discussed * Email: [email protected] 1 Cocchi et al. 2 Criticality in the brain List of abbreviations ATP: Adenosine Triphosphate BS: Burst suppression DC: Direct current DFA: Detrended fluctuation analysis ECoG: Electrocorticography EEG: Electroencephalogram fMRI: Functional magnetic resonance imaging GABA: Gamma-Aminobutyric acid IBI: Inter-burst interval MEG: Magnetoencephalography REM: Rapid eye movement sleep SOC: Self-organised criticality SWS: Slow-wave sleep Tc: Critical temperature 2 Criticality in the brain Cocchi et al. 3 Table of Contents 1. Introduction ......................................................................................................................................... 4 2. Criticality in physical systems ............................................................................................................ 5 2.2. Criticality and phase transitions ...................................................................................... 8 2.3. The conceptual appeal of criticality .............................................................................. 11 2.4. Self-organised criticality ............................................................................................... 12 3. Criticality in the brain ....................................................................................................................... 13 3.1. Rhythmic fluctuations, bifurcations and slowing down ................................................ 13 3.2. Neuronal avalanches and phase transitions ................................................................... 14 3.3. Computational aspects of neuronal criticality ............................................................... 16 3.4. Self-organised neuronal criticality ................................................................................ 18 4. Challenges and pitfalls of the criticality hypothesis ......................................................................... 19 5. Emerging role of criticality in cognition ........................................................................................... 23 5.1. Criticality in brain and behaviour.................................................................................. 23 5.2. Suppression of criticality during task performance....................................................... 24 6. Criticality in disease .......................................................................................................................... 25 6.1. Bifurcations and seizures .............................................................................................. 25 6.2. Crackling noise and neonatal burst-suppression ........................................................... 26 6.3. Criticality and neuropsychiatric disorders..................................................................... 27 7. Summary ........................................................................................................................................... 28 8. References ......................................................................................................................................... 30 Box 1: Ten-Point Summary .................................................................................................................. 42 Box 2: Glossary..................................................................................................................................... 43 3 Criticality in the brain Cocchi et al. 4 1. Introduction Enormous strides have been achieved in neuroscience across a hierarchy of scales of enquiry, from the variety of neural cell types and their molecular biology, through the function of cortical circuits and, in recent years, to the complex architecture of large-scale brain networks. Much of this success has been achieved within research silos, with a focus on scale-specific phenomena, partly mandated by the apertures of various imaging technologies and partly by the training and cultures within the various neuroscientific disciplines. Research in neuroscience also proceeds within a largely descriptive world-view, with increasing emphasis on the collation and statistical characterization of "big data" (Biswal et al., 2010; Markram et al., 2015). Whilst specific mechanisms have been elucidated across an array of basic and clinical neuroscience domains, important challenges remain to be addressed: First, since correlations between behaviour and neuronal activity have been documented at almost every scale of analysis, it seems unlikely that a description of the brain at any particular scale will be sufficient to describe brain function. How is neural activity integrated across scales to give rise to cognitive function? What are the mechanisms linking activity across scales? Second, brain function does not only rely upon the execution of particular functions, but also on adaptive switching from one function to another, depending on context and goals. What are the fundamental principles underlying such complex, flexible neuronal dynamics? Third, what are the major theoretical frameworks to explain and unify the properties of all the large volumes of data currently being accrued? Fourth, how is information encoded by neurons – in the entropy of individual spikes, or via likelihood functions encoded by the distributions of population activity? The principles that unify brain function across spatial and temporal scales remain largely unknown. However, comparable multi-scale challenges exist in other scientific disciplines. Meteorology, for example, spans scales from local wind gusts through regional weather systems up to global climate patterns. Each scale is nested within a larger scale, such that the local variance in wind gusts depends upon the regional weather, which is likewise constrained by global trends such as El Niño. Mathematicians and physicists have developed a considerable armoury of analytic tools to address multi-scale dynamics in a host of physical, biological and chemical systems (Bak et al., 1987). Chief amongst these is the notion of criticality, an umbrella term that denotes the behaviour of a system perched between order (such as slow, laminar fluid flow) and disorder [such as the turbulence of a fast-flowing fluid, (Shih et al., 2015)]. A critical system shows scale-free fluctuations that stretch from the smallest to the largest scale, and which may spontaneously jump between different spatiotemporal patterns. Despite their apparent random nature, the fluctuations in these systems are highly structured, obeying deep physical principles that show commonality from one system to the other (so-called universality). They can hence be subject to robust statistical analysis and modelling. Critical systems thus display the type of cross-scale effects and dynamic instabilities linking activity at different scales that is typical of brain functioning. An emerging literature suggests that brain function may be supported by critical neural dynamics, with original research that continues to flourish (Deco and Jirsa, 2012; Kelso et al., 1992; Priesemann et al., 2014; Scott et al., 2014) on the background of an existing body of reviews and syntheses (Beggs and Timme, 2012; Chialvo, 2010; Deco and Jirsa, 2012; He, 2013; Hesse and Gross, 2015; Kelso et al., 1992; Plenz and Thiagarajan, 2007; Priesemann et al., 2014; Schuster et al., 2014; Scott et al., 2014; Shew and Plenz, 2013). The principles supporting the emergence of these patterns of activity are not yet fully understood but recent studies using neuroimaging techniques such as functional magnetic resonance imaging (fMRI) and electroencephalogram (EEG) (Deco et al., 2009; Linkenkaer-Hansen et al., 2001; Stam and de Bruin, 2004) have added to earlier work in slice preparations (Beggs and Plenz 2003). Computational 4 Criticality in the brain Cocchi et al. 5 models also show that neural systems have maximum adaptability to accommodate incoming processing demands when they are close to a critical point (Friston et al., 2012b; Friston, 2000; Gollo and Breakspear, 2014; Kastner et al., 2015; Shew et al., 2009; Yang et al., 2012). Conversely, brain disorders, as diverse as epilepsy, encephalopathy, bipolar disorder and schizophrenia may correspond to excursions from such an optimal critical point. Despite the ubiquity of criticality in many branches of science, its application to neuroscience is relatively recent and unknown to many neuroscientists. When it is used, it is often invoked metaphorically; a practice which risks mixing distinct processes incorrectly into a rubric term. Research into criticality has much to offer neuroscientists, but needs to be used in accordance with its well-defined operational criteria. Accumulating evidence should also be viewed cautiously according to emerging pitfalls. Here, we first revisit the core notion of critical phenomenon and provide examples from the physical sciences. We then review the classic and recent studies of neuronal criticality. We finally consider emerging applications that advance new theories of healthy and maladaptive cognition using the innovative tools that criticality provides. 2. Criticality in physical systems Criticality refers to the appearance of erratic fluctuations in a dynamical system that is close to losing dynamic stability. Because the nature of the instability can vary (as we review below), criticality is a broad umbrella term that subsumes several related phenomena but also excludes others. In this section, we present a brief pedagogical account of criticality. We first consider critical fluctuations that occur close to instability in systems consisting of only a few interacting components. This allows us to introduce core signatures of criticality; namely the emergence of scale-free temporal fluctuations, slowing down and multistability, defined below. We then consider criticality in complex systems composed of many interacting parts. These extra degrees of freedom allow for the occurrence of scale-free spatiotemporal fluctuations known as avalanches. Finally, we consider self-organized criticality - that is, the process by which criticality emerges without the need for external tuning of a control parameter. 2.1. Criticality and bifurcations We first consider dynamical systems composed of only a few interacting components. Consider the classic example in which two species interact as predator and prey. When the interactions amongst the species are weak and alternative food sources are available, relatively simple models predict that the populations of both species reach stable equilibria, and the processes of consumption and reproduction occur at a steady rate (Berryman, 1992). However, if the interactions between the species increase (i.e. the predators rely more heavily upon the prey population), there reaches a critical point of interactivity above which the two populations begin to oscillate: When the predator population is relatively low, the number of prey animals grows through unbalanced reproduction. This then yields a ready food source for the surviving predators, whose population increases. However, as the increasing numbers of predators consume the available prey species, the numbers of the latter then decline, with a subsequent effect on the survival of the predators; the cycle then begins anew. This transition from steady state to cyclic behaviour due to strong interactions is called a bifurcation (Figure 1a). The strength of the interaction is called a control parameter and the point at which the bifurcation occurs is denoted the critical point. For three or more interacting species, further bifurcations to more complex dynamics can occur, leading from periodic to chaotic oscillations (smooth and deterministic but aperiodic oscillations) (Arneodo et al., 1980; Vano et al., 2006). 5 Criticality in the brain Cocchi et al. 6 There are two crucial variations on this simple example. First physical processes, such as the predator- prey example, inevitably occur in the presence of small but unceasing random fluctuations. This noise arises from a myriad of causes such as the probabilistic nature of individual predator-prey encounters as well as influences not explicitly modelled (diseases, fluctuating environmental conditions etc). When the interactions between predator and prey are weak, the equilibrium state is very stable and the presence of such random fluctuations only have a minor impact on the observed steady state populations. Likewise, if the interactions are strong, the cyclic oscillations in population numbers are also very stable: The amplitude of the oscillations is relatively stable and the noise is again effectively suppressed. More technically, away from the critical point, the system is said to be dynamically stable and the fluctuations are strongly damped, dropping off quickly (with an exponential decay rate). However, in the immediate vicinity of the critical point, the perturbations grow in magnitude, dominating the observations because the system is less stable (or weakly stable). That is, the variance of the observed fluctuations grows in magnitude. Moreover, the fluctuations decay slowly. To be precise, the decay of fluctuations in time changes from a fast (exponential, Figure 1b, inset) to a slow (power-law, Figure 1c, inset) process. Fluctuations with power-law correlations are scale-free because they do not have a characteristic time scale. Upon further increases in the control parameter, the fluctuations in the envelope of the oscillations quickly become stable (Figure 1d) and drop off quickly (exponentially, likewise the inset of Figure 1b). Hence near the critical point of a bifurcation, we encounter two central features of criticality: High amplitude scale-free fluctuations (Figure 1h) and slowing down (i.e., longer autocorrelation, Figure 1i). These large, slow fluctuations are termed crackling noise after the sound they make if played audibly (Sethna et al., 2001). A second variation concerns the nature of the bifurcation itself. The predator-prey model exhibits a classic bifurcation, whereby a single critical point separates two distinct behaviours (steady state and oscillatory) in parameter space. This is denoted a supercritical bifurcation because the cyclic oscillations occur for values of the control parameter strictly greater than the critical point. However, subcritical bifurcations are also possible. In this setting, there exists a region where the steady state solutions and the periodic oscillations co-exist. Outside of this zone, the system behaves in the same way as the supercritical bifurcation (i.e., a single steady state or a periodic pattern of activity) (Figure 1e). However, within this zone, the two dynamic states co-exist (Figure 1e and f). Noise can then push the system between these stable states, causing erratic jumps between low amplitude equilibrium and high amplitude oscillations (Figure 1f). This type of behaviour is called multistability (Freyer et al., 2011; Tognoli and Kelso, 2014). We consider examples of multistability in brain and behaviour below. Although critical fluctuations (near a supercritical bifurcation) and multistable fluctuations (due to a subcritical bifurcation) are mathematically related, they yield quite distinct statistics: The former (slow fluctuations) follow a scale-free power law distribution. This is the classic meaning of the term criticality. Noise-driven switches between two or more multistable attractors do not occur with a scale-free probability. In the setting of large additive noise, the transitions are akin to a Poisson process and thus follow an exponential distribution. With smaller, state-dependent noise, the system tends to get trapped near each state, with the transitions then following a heavier tailed stretched exponential distribution (Freyer et al., 2011; Freyer et al., 2012). Either way, such multistable switching does not possess scale-free (power law) properties and does not correspond to the classic notion of criticality. 6 Criticality in the brain Cocchi et al. 7 Figure 1: Criticality in a low dimensional system consisting of a few interacting components. (a) Super-critical bifurcation diagram, depicting the amplitude of a system's state variable (y-axis) as a function of a control parameter (such as the strength of interactions, x-axis). When the control parameter is increased, the activity of the system switches from a damped equilibrium point (red circle) to oscillatory behaviour (yellow circle). The point of change is known as the critical point (blue circle). (b) In the presence of noise, the damped system (red circle) exhibits low amplitude, rapid fluctuations. The duration of these follows an exponential probability distribution (red dots, inset). (c) At the critical point, the fluctuations have high variance and rise and fall slowly, following a power law distribution, corresponding to a linear relationship between their duration and their likelihood in double logarithmic coordinates (inset). In addition, the slope of this relation is described by a critical exponent of α = −3/2. (d) Beyond the critical point, the system exhibits sustained oscillations. The fluctuations in the amplitude envelope of the oscillations are fast and small. (e) Sub-critical bifurcation 7 Criticality in the brain Cocchi et al. 8 diagram, with a zone of co-existence (or "bistability") between the fixed point and oscillatory behaviours. In this case, system noise not only drives fluctuations around each attractor, but can also drive sudden and erratic jumps between the two dynamic states as depicted by the double headed arrow. (f) Example bistable time series. (g) Multistable bifurcation diagram as reproduced from Freeman (Freeman, 1987), proposed as a model for perceptual activity in the olfactory system. In this case, the number and complexity of the attractors is larger, however, the underlying principle is the same. Critical slowing down corresponds to a sharp increase in the coefficient of variation of the mean amplitude across 200 trials (h) and the auto-correlation function (i) at the critical point. 2.2. Criticality and phase transitions We have thus far considered relatively simple systems composed of only a few components, or where the elements of the system are lumped into a small number of variables (such as all predators and all prey species each being considered a single entity). We now move to studying critical systems composed of many interacting components such as magnetic spins in iron (Stanley, 1987), grains of sand falling onto a pile (Bak et al., 1988), or neurons (Plenz and Thiagarajan, 2007). On top of the example considered in Figure 1, these examples introduce a spatial dimension through which the components of the system interact. The emergence of a magnetic field in a ferromagnetic material (such as iron) cooled below a critical temperature (Tc) is a classic example of a phase transition. Such materials have permanent magnetic moments (dipoles) due to the spin of unpaired electrons in atomic or molecular electron orbits. These dipoles interact through the mutual effects of the local fields that they impart on their immediate neighbours (Figure 2a), causing neighbouring dipoles to align and form local domains of coherent fields. These effects partially counter the influence of stochastic thermal and quantum effects that cause random flips in the direction of the dipoles. At temperatures greater than Tc, stochastic flips disrupt the formation of larger domains and, in the absence of an external field, the material will not possess a macroscopic magnetic field. Slow cooling of the material allows domains of increasing size to form, although domains at the very largest scales continue to disappear into the background noise. However, when the material is cooled to a critical point (the Curie temperature), the smaller domains coalesce into increasingly larger ones until they approach the size of the entire system. The coalescence of small domains into those of successively larger size is called an avalanche (Figure 2a). At the Curie temperature, avalanches have no characteristic size and thus may intermittently sweep through the entire system. These avalanches can be measured empirically using a large, external pick-up device (Cote and Meisel, 1991; McClure Jr and Schroder, 1976; Meisel and Cote, 1992; Perković et al., 1995). Below the critical temperature, the mutual interactions amongst the spins align into domains that encompass nearly every dipole; at this point, the material shows a coherent ferromagnetic field (despite ongoing disorder at the atomic scales). The transition through the critical point in such a high dimensional system is called a phase transition. A phase transition in iron cooled below its Curie temperature is a classic example of how simple internal interactions can overcome disorder and yield, through a critical point, a macroscopic field. The imposition of an external field of sufficient strength on a ferromagnet below the critical point can cause the field to suddenly switch directions to align with the applied field (Vojta et al., 2013): Here, in contrast, the macroscopic order is imposed externally. There are many similarities between bifurcations and phase transition, including the presence of super- and subcritical varieties (Figure 2b): These are called continuous and discontinuous or (second- and first-order) phase transitions in this context (Kim et al., 1997). The transition from a para- to a ferromagnet due to cooling, outlined above, is an example of a continuous phase transition: 8 Criticality in the brain Cocchi et al. 9 The sudden switching of that field due to an external field is a discontinuous one, as is water turning into vapour in the presence of heat (Stanley 1987). When the phase transition is discontinuous, then noise-driven multistability may also occur: Noise can induce switches between ordered and random states (Figure 2b). Alternatively, as in the case of the (discontinuous) phase transition between water and steam, there can arise complex mixtures of both. Just as in the case of a bifurcation, however, multistability arising due to a discontinuous phase transition does not exhibit the scale-free, 'critical phenomena' discussed above. Critical, power law scaling in the spatial and temporal domains is unique to a continuous phase transition. Whilst we here focused upon the canonical example of spins in a weak external field, the basic ingredients (local interactions, noise, a large number of subsystems, an external influence that brings weak coherence) occur widely and, as a result, phase transitions are ubiquitously observed in nature (Stanley, 1987). The temporal behaviour of a spatially-extended system near a phase transition mirrors the behaviour of two interacting elements near a bifurcation – namely slow, high amplitude fluctuations. These fluctuations, however, additionally exhibit complex spatiotemporal processes – avalanches – that also show scale-free statistical properties. Away from the critical point, the likelihood of an avalanche drops off quickly (exponentially) with size (the number of elements involved). In the classical example of magnetism, the domains of coherent spins are very small at high temperatures. If the temperature is tuned towards the critical point (Tc), the coherent domains sporadically increase in size and the ensuing distribution of avalanche sizes decays slowly as a function of spatial scale. The size of such domains measured over time converges toward a scale-free (power law) effect (Figure 2c). If the temperature is further reduced, large and stable coherent domains appear corresponding to the emergence of an internal magnetic field. The correlation length is a useful concept in this setting. In the disordered phase of the system (i.e. for weak external fields) only adjacent spins are correlated– distant spins are completely uncorrelated. As larger avalanches begin to appear at low temperature, electron spins become correlated across the scale of the corresponding coherent domains. The correlation length – the spatial scale at which pairs of electrons are at least weakly correlated –increases. At the phase transition, as the size of the coherent domains approaches that of the system, the correlation length diverges upwards. As a result, external perturbations applied to any part of the system may lead to a change in the state of the whole system. Put alternatively the system has maximum dynamic range because any small perturbation has a chance of changing the electron spins of such a critical paramagnetic system. We have focused upon phase transitions in spatially extended (embedded) systems, such as magnets and water which are dominated by interactions or collisions between the neighbouring elements of the system. However, the description of avalanches in large N systems does not inevitably refer to space. A branching process is a simple model of a phase transition that describes how activated elements may either decay (to inactive) or activate other elements with the progression of time (De Carvalho and Prado, 2000). Criticality occurs when the decay and activation rates are in equal ratio. This canonical model of a phase transition, which has been used to study criticality in neural systems [e.g. (Beggs and Plenz, 2003)], does not require a spatial dimension; metrics such as the correlation length do not make sense in these models. Although such abstract models may not take space into account, complex multi-unit physical systems such as the brain must be embedded in space, and the interactions between the elements are very often constrained by their spatial proximity (Roberts et al., 2016). 9 Criticality in the brain Cocchi et al. 10 Figure 2: Phase transitions and avalanches in spatially extended systems. (a) Continuous (or second order) phase transition. The disordered (random) phase (red) shows lack of spatial order (red square) with randomly oriented spins in a typical physical system such as a ferromagnet in a weak external field. The ordered phase (yellow) shows large domains of co-aligned spins. At the critical point (blue), avalanches of complex partially- ordered domains rise and dissolve across all scales (blue square), leading to a power law size distribution. (b) Phase transitions can also be discontinuous (also called first order). As with a sub-critical bifurcation, system noise then causes erratic switching between the disordered and ordered phases (double headed arrow). (c) Cumulative probability distribution of the relationship between the size and likelihood of avalanches at criticality. A scale-free processes yields a linear scaling relationship in double logarithmic coordinates (a power law) with a critical exponent of α = −3/2. Note the slight exponential truncation at the right hand side, due to finite size effects. As we have seen, phase transitions can be considered a natural extension of the notion of a bifurcation from systems with few components, to those with many. The underlying mathematics is very similar 10 Criticality in the brain Cocchi et al. 11 (in fact, the so-called Landau equation, used to model generic phase transitions, is mathematically identical to the Normal form equation used to describe bifurcations). Historically, however, the two phenomena have been studied in different fields - bifurcations by applied mathematicians, whereas phase transitions were classically the domain of physicists. This legacy has led to a difference in the use of the central terms, sub- and super-criticality: In mathematics, "sub-" and "super-critical" qualify bifurcations, denoting distinct instabilities that differ in their underlying mathematical nature (Figure 1a and 1e) and, as a result, the behaviour they yield. In physics, these terms are used to describe the phases of the system. "Subcritical" phase is used to denote the stable, absorbing state below a phase transition (Figure 1a, red dot and Figure 1b). The term "supercritical" is used to denote the ordered state above the transition, whereby the amplitude of the order parameter (y-axis) is typically non-zero (Figure 1a, yellow dot and Figure 1d). It is unlikely that an attempted synthesis of those terms here would pervade both fields. For the remainder of this paper, we use the terms "sub-" versus "super- critical bifurcation" to denote the type of instability, and "sub-" versus "super-critical state", "activity" or "phase" to denote where a particular system lies with respects to the phase transition. In general, since the interpretation is largely dependent on the audience, caution is required when interpreting or using those terms more broadly. Phase transitions and criticality have been documented in a very broad range of physical systems over many decades (Kosterlitz and Thouless, 1973; Yang and Lee, 1952) and their study remains one of the most active areas of research in branches of physics such as statistical mechanics (Papanikolaou et al., 2011; Sethna et al., 2001; Zapperi et al., 2005). The signatures of criticality have been documented in systems as diverse as flocking birds (Cavagna et al., 2010), earthquakes (Burridge and Knopoff, 1967; Carlson and Langer, 1989; Rice and Ruina, 1983); solar flares (Lu et al., 1993), armed conflict (Roberts and Turcotte, 1998), traffic jams (Nagel and Herrmann, 1993), and capital wealth (Roberts and Turcotte, 1998) – even crumpled paper [(Houle and Sethna, 1996; Kramer and Lobkovsky, 1996); for review see (Roberts and Turcotte, 1998)]. We turn to evidence for criticality in the brain in Section 3 after considering its underlying appeal. But before moving, it is instructive to contextualize the importance of criticality as a theoretical framework and how it may come about in many distinct systems. 2.3. The conceptual appeal of criticality Criticality derives its basic appeal from a number of considerations. First, it speaks to the presence of a relatively simple underlying process - the response of a weakly stable system to stochastic perturbations - arising in very different settings. The processes that drive the system close to instability can be diverse – a build-up of fuel; varying temperature; a driving external magnetic field; strong interactions between species – but the collective response in generating slow, multiscale fluctuations is shared. Likewise, the many specific details of the systems differ markedly (e.g., magnetic spins, moving tectonic plates, neurons) but can be unified by their core dynamic nature - possessing interactions among their components that erratically amplify and damp microscopic perturbations. This notion of "universality" is very appealing to scientists who seek unifying principles across diverse systems. That is, the appearance of power-law and invariant scaling in markedly different systems suggests the importance of processes that transcend their particular incarnation. Of note, the characteristic exponent in the power-law scaling that describe the critical fluctuations in many of these systems typically converges to a value of −3/2. Theoretical considerations support the emergence of this value in systems at the cusp of a phase transition (Zapperi et al., 1995). Thus, basic theoretical arguments unify diverse phenomena – this is the essence of universality (Stanley, 1999). 11 Criticality in the brain Cocchi et al. 12 Computational considerations also underlie the appeal of criticality. Here we review the computational aspects of criticality in physical systems. In Section 3.3, we focus on the computational advantages of criticality in the brain. The earliest demonstrations of the computational advantages of the critical state were done in a very simple and idealized system called cellular automata (CA), whose dynamics evolve discretely in space and time according to very simple interaction rules (Langton, 1990). By changing an underlying interaction parameter, CA can be tuned to converge very quickly to simple periodic (spatiotemporal) structures, or to unstructured, chaotic processes. In between these scenarios – at the so-called "edge of chaos" – CA exhibit lengthy mixtures of ordered and disorganized structures. Theoretical arguments show that the computational complexity of CA diverge in this regime. That is, if one considers the information content of the system at each time point, the number of iterations before CA converge onto a stable solution becomes very long for this in-between state (although see (Mitchell et al., 1993) for an opposing position). Depending upon one's viewpoint, the simplicity of CA is either conceptually appealing (since complexity arises from very simple laws) or distracting, because the physical meaning of CA is unclear. However, the implications of the proposal – complexity from simplicity - are tantalising, underlying the influence of CA. Its catch-phrase "edge of chaos" became a very well-known way to refer to the computational advantages offered by the critical state. Further research has shown that several other physical and biological systems also have optimal computational properties at criticality (Crutchfield and Young, 1988; Kauffman and Johnsen, 1991; Mora and Bialek, 2011; Nykter et al., 2008). The appeal of criticality also finds support from thermodynamic perspectives. In a stable, subcritical system, random fluctuations arising from thermal energy and other sources of entropy are confined to the microscopic scale, like a giant TV screen showing pixel-wise static. While these microscopic fluctuations have high entropy, meso- and macroscopic scales are damped and are hence in a featureless, low entropy state. Above the critical point, the macroscopic scale of the system can show interesting features, such as periodic structures and oscillations. However, fluctuations at finer scales are slaved to these large-scale features and thus do not express the potential entropy arising from the smallest microscopic scales. In these two states – sub- and super-critical, respectively, high entropy can be thought of as being trapped at one particular scale and unavailable at other scales. At the critical state, microscopic fluctuations erratically disseminate to larger scales through avalanches and crackles. These fluctuations introduce packets of disorder which accordingly increase the information content of the system across all scales. That is, while criticality increases correlations – and thus decreases entropy – at the smallest scale, it "transports" random fluctuations across scales increasing the total complexity of the system (Tononi et al., 1994). 2.4. Self-organised criticality Why is it that so many systems found in nature appear to be perched at a critical point? In theory the critical point becomes confined to a very small range of values as the size of the system increases (see Figure 1h). For an experimental system such as the paramagnetic material discussed above, the external field can be carefully (manually) tuned until crackling noise and avalanches appear. However, for other systems, such as earthquakes, forest fires and flocking birds the fingerprints of criticality seem to arise internally without the need for careful tuning by an external observer. The answer to this apparent dilemma is contained in the notion of self-organised criticality (SOC), a process whereby a complex system is driven toward its critical point across a very wide set of starting points and parameter values. The classic example of SOC was provided in the behaviour of sand-pile avalanches by the work of Per Bak (Bak, 1990; Bak et al., 1987, 1988). In essence, the slow addition 12 Criticality in the brain Cocchi et al. 13 of sand to the apex of a sand-pile leads to the gradual increase in the slope of its sides. At a critical slope, scale-free avalanches of falling sand begin to occur. The slope angle decreases with each avalanche as (gravitational) energy is released from the system, then increases again as new sand is added. The slope does not need to be tuned by the experimentalist but naturally emerges from the interplay of the interactions between the adjacent grains of sand, the external (gravitational) force and the slow addition of sand. Modelling SOC in complex systems with weak local interactions and noise is a very active field (Marković and Gros, 2014). Two mechanisms appear sufficient for the appearance of SOC – firstly the dissipation of energy and secondly some form of "memory" in the system. For example, a large forest fire burns through a build-up of timber fuel: Energy has dissipated from the system and a period of time must now pass until there is sufficient new fuel for fire of any appreciable magnitude. Likewise, a large avalanche of sand in a slowly building sand pile changes the gradient of the pile. Time and small avalanches must then accrue before the slope of the pile is sufficiently steep to trigger another large avalanche. Again in seismology, tension from tectonic plates is released following a large earthquake and its aftershocks, such that subsequent large earthquakes are unlikely to follow immediately. Each of these systems is characterized by the build-up and subsequent dissipation of energy or resources, whose release is 'remembered' by the system until sufficient resources have recovered. We will encounter similar concepts when we discuss mechanisms of SOC in neuronal systems (Section 3.3). 3. Criticality in the brain The role of criticality and multistability in neurophysiological systems of the brain was first articulated over 3 decades ago by Walter Freeman following detailed empirical analyses and computational models of the rabbit olfactory bulb (Figure 1g). In particular, Freeman proposed that the process of inhalation and exhalation acted, via modulation of the gain of excitatory neurons, to sweep the activity of the olfactory bulb through a sub-critical bifurcation and hence through a zone of multistability (Freeman, 1987; Freeman, 1991). Sensory inputs, arising from contact of inhaled molecules with membrane receptors of olfactory neurons, then act to selectively perturb the system onto one of several competing dynamic patterns. This dynamic pattern was proposed to encode the olfactory input – the percept - until it destabilized during exhalation as the system passed again into the zone where only the stable equilibrium solution exists. We now survey more recent examples of criticality and multistability in neuronal systems that build upon the foresights contained in Freeman's prescient papers. 3.1. Rhythmic fluctuations, bifurcations and slowing down The mechanisms underlying motor coordination have been an intriguing area for the application of dynamic systems theory (Bressler and Kelso, 2001; Kelso and Clark, 1982). One fruitful candidate has been the study of rhythmic finger tapping. At slow frequencies, humans are able to tap in either of two stable modes: syncopation and anti-syncopation. However, at high frequencies, the anti- syncopation mode becomes unstable and only the in-phase syncopation pattern is expressed (Kelso et al., 1986). The hallmarks of criticality are seen just prior to this transition, namely slowing, high amplitude movement fluctuations (Kelso, 1984, 2014). To explain this, Haken, Kelso and Bunz used a sub-critical bifurcation in a simple model of motor coordination between the left and right motor cortices (Haken et al., 1985). The transition from a bimodal to unimodal pattern of behaviour occurred at the critical value of the movement frequency. This framework has also been employed to 13 Criticality in the brain Cocchi et al. 14 explain the transition between movement patterns induced by transcranial magnetic stimulation (TMS) (Kelso, 2014). More recent work on criticality has focused on the temporal fluctuations observed in the major rhythms of EEG and MEG data. Employing an analysis called detrended fluctuation analysis (DFA), Linkenkaer-Hauser and colleagues reported that the fluctuating amplitudes of the two dominant oscillations of the human brain – the alpha and beta rhythms – exhibited scale-free temporal statistics (Linkenkaer-Hansen et al., 2001). Fluctuating levels of synchrony between pairs of electrodes have also been reported to show scale-free statistics (Stam and de Bruin, 2004). Likewise, the power spectrum of human neocortical activity acquired from invasive ECoG data shows scale-free temporal behaviour across a very broad range of frequencies (Miller et al., 2009). These frequencies also show multi-scale nesting – that is, the amplitude of high frequencies is coupled to the phase of lower frequencies; a pattern that is recursively repeated from very slow to very high frequencies (He et al., 2010). Computational models of large-scale neuronal activity – neural field models – suggest that the critical temporal statistics in these electrocortical recordings may arise from a subcritical bifurcation of activity in corticothalamic loops (Freyer et al., 2009; Freyer et al., 2011). Such modelling proposes that noise-driven switching between a low amplitude steady state and high amplitude oscillations (Figure 1) yields the empirically observed critical fluctuations seen at rest (Freyer et al., 2012). 3.2. Neuronal avalanches and phase transitions In 2003, Beggs and Plenz found evidence of critical behaviour in the erratic spontaneous activity measured in in vitro neuronal cultures (Beggs and Plenz, 2003). They documented the two salient features of criticality in a spatially extended critical system, namely power-law scaling in time (the duration of bursts of activity) and space (the number of electrodes spanned by each burst). Together with the earlier work of Freeman, this finding ushered in criticality as a term of clear relevance to complex neuronal multi-scale phenomenon. Since these initial reports of avalanches in in vitro slice preparations by Beggs and Plenz, research into critical avalanche-like activity in spatiotemporal neural recordings has proceeded at great pace (Plenz and Thiagarajan, 2007; Schuster et al., 2014; Shew, 2015). Observations of scale-free spatiotemporal fluctuations in spontaneous, physiological data have progressed from in vitro slice preparations (Beggs and Plenz, 2003), to in vivo recordings from superficial layers of cortex (Gautam et al., 2015; Gireesh and Plenz, 2008) to awake non-human primates (Petermann et al., 2009) [for review, see Shew and Plenz (2013)]. Scale-free avalanches have been reported in human whole brain magnetoencephalographic (MEG) data (Shriki et al., 2013), and complex, scale-free spatial dependences, consistent with avalanches, have been described in whole brain functional neuroimaging (fMRI) data (Tagliazucchi et al., 2012). Notably, these recordings cross broad scales of aperture from multi-unit recordings to macroscopic field potentials and whole brain functional neuroimaging data. There is a growing focus in the imaging community on spontaneous (resting-state) fMRI data and the reproducible structures these reveal in health (Damoiseaux et al., 2006; Zalesky et al., 2014) and disease (Fornito et al., 2015). Whereas the number of channels in neurophysiological recordings, such as MEG, has a modest upper bound (of several hundred), the high spatial resolution of fMRI yields thousands of voxels (~100,000 voxels). These data hence contain the breadth of spatial scales required to interrogate whether the spatial fluctuations are scale-free (Eguiluz et al., 2005) and thus whether critical dynamics underlie the dynamic patterns seen at rest (Chialvo, 2012). Recent evidence from both empirical (Tagliazucchi et al., 2012) and computational (Deco and Jirsa, 2012) analyses points in 14 Criticality in the brain Cocchi et al. 15 favour of this proposal. Among the most intriguing findings are the recapitulation of the classic resting-state networks (Yeo et al., 2011) by models of critical dynamics arising from primate (Honey et al., 2007) and human (Haimovici et al., 2013) structural connectomes. Analysis of the temporal statistics of resting-state fMRI and EEG also suggests that long-range, scale-free correlations may indeed lie at the very heart of the slow fluctuations that are observed in these data (Van de Ville et al., 2010). This emerging view is schematically summarized in Figure 3. In the sub-critical zone, bursts of cortical activity are sporadic uncoordinated (red box). Above the critical value, cortical activity is coupled too tightly and conversely, inadequately segregated (yellow box). Resting-state networks function at the critical value, where switching between network states occurs due to weak dynamic instabilities (light blue box). Figure 3: Proposed role of criticality in large-scale, resting-state brain dynamics. In the sub-critical region, individual brain regions are effectively uncoupled, showing a lack of integration (red square). Conversely, in the super-critical region, integration is too great and there is a lack of segregation (yellow). Near the critical point, an emerging body of work in EEG, MEG and fMRI suggests that brain systems show a dynamic balance of integration and segregation (blue square), fluctuating among the various resting-state networks (and EEG rhythms). Cognitive function requires a slight incursion away from the critical regime leading to a stabilization of one particular network, consistent with the earlier proposals of Freeman. There is now a very well established relationship between resting state brain networks and the underlying structural connectome from which they arise (Honey et al., 2007). Criticality arising in simple systems does not require a complex spatial substrate: Rather as we have reviewed above, its hallmark is the emergence of complex spatiotemporal processes from simple, local interactions. However, complex networks may allow critical-like behaviour to occur in a region of parameter space instead of a single point (Moretti and Muñoz, 2013). Moreover, in the primate brain, structural- functional correlations argue for the existence of a relationship between critical dynamics and the relatively static underlying structural connectome. The nature of this relationship between critical states in resting-state fMRI data and the connectome is not well understood. Highly interconnected 15 Criticality in the brain Cocchi et al. 16 cortical hubs, and brain regions comprising the so-called default-mode network may play a prominent role in maintaining resting-state network dynamics (Gollo et al., 2015; Leech et al., 2012; Vasa et al., 2015) and in facilitating the efficient spread of avalanche events through macroscopic brain networks (Misic et al., 2015). The constellation of densely connected hub regions – the rich club – appear to support a slow, stable dynamic "core" whereas peripheral sensory regions introduce (rapid) stimuli- related variability in the system (Bassett et al., 2013; Gollo et al., 2015; Hasson et al., 2015). Such a core-periphery organization of brain network dynamics speaks to a hierarchy of time-scale fluctuations, in which hub regions integrate and regulate the network dynamics largely operating at slow frequencies (Cocchi et al., 2016; Gollo et al., 2017; Gollo et al., 2015; Hasson et al., 2008; Honey et al., 2012; Murray et al., 2014). Hub regions within the default-mode brain network may represent a structural signature of near-critical behaviour. Regions comprising this network exhibit coordinated activity in the resting-state when the coherence between nodes of other "task-positive" networks is generally suppressed (Cocchi et al., 2013; Fox et al., 2005; Hearne et al., 2015). These observations suggest mechanisms through which critical dynamics may adapt to, and reshape the complex nervous systems in which they occur, particularly the relationship between synaptic processes, functional connectivity and network topology (Rubinov et al., 2009; Zhigalov et al., 2017). 3.3. Computational aspects of neuronal criticality The pioneering work of Freeman (on bifurcations and multistability), and Beggs and Plenz (on critical avalanches) provided proof-of-principles that the science of criticality could be used to inform our understanding of complex patterns of activity in the brain and, by extension, behaviour. Research in these areas has accelerated dramatically and now yields a stream of important discoveries spanning from the neuronal (Gal and Marom, 2013; Gollo et al., 2013) to the whole-brain scale (Kitzbichler et al., 2009). As reviewed above, the study of criticality in physical systems using simple models suggested that systems at the critical state are endowed with optimal computational properties. Such advantages have recently been demonstrated in models and empirical recordings of critical neuronal systems [for review, see (Beggs, 2007; Shew and Plenz, 2013)]. Perhaps most crucially, optimal dynamic range – the sensitivity of a neuronal system to respond to, and amplify, inputs across a broad spectrum of intensities – was shown to be maximized in models of neuronal systems tuned to a critical state (Kinouchi and Copelli, 2006; Larremore et al., 2011). In slice cultures grown on the surface of multielectrode arrays, Shew and colleagues later provided empirical evidence for this proposal, showing that the maximum dynamic range to electrical perturbation was indeed maximized when the cultures were pharmacologically manipulated to be close to criticality (Shew et al., 2009). Recent electrophysiological recordings from the anaesthetized rat provided the first in vivo evidence that dynamic range in perceptual systems is maximized when background activity is at the critical point (Gautam et al., 2015). Another example of the computational advantages of criticality arises from simplified neuronal models which predict high fidelity and optimal information transmission (maximum mutual information between sender and receiver) at criticality (Beggs and Plenz, 2003; Greenfield and Lecar, 2001): This prediction was also later observed in vitro (Shew et al., 2011) and more recently in awake, behaving mice (Fagerholm et al., 2016). Research in this field has also suggested that information storage and capacity – the ability of a system to encode a broad repertoire of complex states from which information can be decoded 16 Criticality in the brain Cocchi et al. 17 ⁄ 𝑚) , where Fmax corresponds to the saturated response, and 𝑆0 (Gatlin, 1972)– may be optimized at criticality. Again, this notion has been captured in simple neural models (Bertschinger and Natschläger, 2004; Haldeman and Beggs, 2005; Yang et al., 2017) and also demonstrated in empirical recordings, including those arising in unperturbed (resting state) recordings (Breakspear, 2001; Deco and Jirsa, 2012) as well as through careful pharmacological manipulations of in vivo neurophysiological recordings (Stewart and Plenz, 2006). Selective enhancement of weak (but not strong) stimuli also occurs near a phase transition (Copelli, 2014). It is also interesting to note that optimal computational properties can arise at both continuous and discontinuous phase transitions (Gollo et al., 2012). A compelling argument for the advantage provided by the critical state derives from the analysis of psychophysics experiments, which quantify the relationship between physical stimuli and perceptual responses. Psychophysics relations may represent the earliest documented evidence of critical dynamics in the nervous system (Kello et al., 2010). Across a variety of sensory modalities, these experiments showed power-law relations, known as Steven's laws, in which the perceived psychophysical or neuronal response F is given by: 𝐹(𝑆) ∝ 𝑆𝑚, where S is the stimulus level and m is the Stevens exponent (Stevens, 1975). To account for the saturation of the response that occurs for extreme stimuli, the psychophysics response is modelled by a sigmoid Hill function: 𝐹(𝑆) ∝ the input level for 𝐹𝑚𝑎𝑥 𝑆𝑚 (𝑆𝑚 + 𝑆0 half-maximum response. Importantly, this function also retains the power-law regime governed by the exponent m. This power-law behaviour has a key putative function: It allows animals to distinguish stimulus intensity varying across many orders of magnitude. As such, the power-law regime can compress decades of stimulus variation S into a single decade of response F. A standard means to measure this coding performance is called the dynamic range. The larger the dynamic range, the better the ability to detect changes in stimuli. Modelling the neuronal behaviour at the sensory periphery, Kinouchi and Copelli showed that a large dynamic range emerges from the collective response of a network of many interacting units, and, more importantly, the dynamic range is optimal when the network is at the critical state (Kinouchi and Copelli, 2006). This work explains the long- lasting psychophysical scaling relations (Steven's law), and provides a clear example in which a meaningful biological feature is optimised at criticality. Similar to earlier proposals, the work of Kinouchi and Copelli was also based on a fairly simple model of neuronal activity. Crucially, however, the model generated a prediction that was subsequently verified experimentally (Shew et al., 2009). The explanation of the power-law regime of psychophysics laws in terms of the optimal sensitivity of critical states was an important contribution. However, as usual in analytic approaches, some simplifications were made, leading to at least two conundrums. The first challenge refers to the important trade-off between sensitivity and specificity. The finding that optimal signal coding occurs at criticality implies maximum sensitivity. Yet, the specificity of this state is compromised because of increased levels of fluctuation. As illustrated in Figures 1 and 2, criticality corresponds to the state with the largest macroscopic fluctuations. This is why the critical state allows for the amplification of stimuli of small intensity, which enhances the ability to distinguish the stimulus intensity varying over orders of magnitude (i.e., large sensitivity). However, the very same effect is also a limitation because the high fluctuations of criticality reduce the specificity of the system. The issue is whether the improved sensitivity remains beneficial to the system when the reduction in specificity is also taken into account. Fortunately, a solution for the sensitivity-specificity conflict exists in the presence of diversity amongst components of the system. Heterogeneous excitable systems exhibit recruitment properties in which units are recruited following their order of excitability (Gollo et al., 2016). Therefore, a state of optimal sensitivity has units forming a subpopulation in a critical regime as well 17 Criticality in the brain Cocchi et al. 18 as subpopulations operating in a non-critical state with improved reliability. Hence, optimised systems represent the coexistence of subpopulations of reliable units and poor sensitivity (poised away from their critical state) with a subpopulation of unreliable units and great sensitivity (typical of the critical state). In other words, optimal perceptual performance may rely on the contribution of critical and non-critical units (Gollo, 2017). The second challenge emerging from the explanation of the power-law regime of psychophysics laws derives from to the fact that the critical state for optimal psychophysical response of Kinouchi and Copelli separates an active state from an inactive state (Kinouchi and Copelli, 2006). The issue here is that the inactive state corresponds to an absorbing state in which the system cannot escape unless external stimuli are provided. Hence, if the system falls into this state, it will get trapped there. Critical systems typically do fall into this state because of their enhanced fluctuations. If the system corresponds to the brain or a part of the nervous system, such an inactive state would be expected to occur rather frequently. However, such silent states are not observed in vivo. The solution to this issue also derives from the incorporation of an essential, but often overlooked, ingredient: inhibition. Larremore and colleagues showed that ceaseless activity and critical avalanches coexist when a substantial fraction of the units are inhibitory (Larremore et al., 2014). Hence, evidence is gradually accumulating for the previously idealised proposal that biological systems can exploit special features of the critical state. Meanwhile, models supporting this proposal are incrementally incorporating greater physiological detail. Research in the auditory system has highlighted the advantages conferred by active responses of hair bundles poised in a critical state (Camalet et al., 2000; Eguíluz et al., 2000). In response to sound- wave input, hair bundles oscillate and their dynamics vary from a steady to an oscillatory regime. Each bundle is tuned to a particular natural frequency and adjacent cells respond maximally to successive pitches, giving rise to a tonotopic map in the cochlea (Romani et al., 1982). Crucially, computational modelling suggests that active and nonlinear hair bundles that operate near a Hopf bifurcation optimise the cochlea's performance and enhance the main features of auditory coding, such as, amplification, frequency selectivity and compressive nonlinearity (Hudspeth, 2008; Maoiléidigh et al., 2012). 3.4. Self-organised neuronal criticality A growing body of empirical work has thus asserted the presence of scale-free statistics across a diversity of in vitro and in vivo neural recordings, while computational models have highlighted its computational advantages. Of note, the critical activity observed by Beggs and Plenz (2003) was stable for many hours and did not require careful tuning of the parameters of their culture (i.e. the pH, temperature, etc). Simple models of criticality classically rely upon fine-tuning of system parameters to a critical value (Levina et al., 2014). What is the basis for a robustness that apparently eschews the need for such a balancing act? As discussed above (Section 2.4), analyses of critical systems in physical systems reconcile this paradox by recourse to self-organised criticality (SOC). In brief, SOC arises when the interactions amongst system components are imbued with some form of plasticity, such as when system energy accumulates and is then dissipated by a large-scale avalanche. During periods of quiescence, energy slowly accumulates until it tips the system into (or above) criticality. The consequent energy dissipation briefly renders the system sub-critical until further energy accumulates. There is thus a time scale separation between the fast system dynamics and the slow build up and dissipation of energy. 18 Criticality in the brain Cocchi et al. 19 Several models of criticality in the brain incorporate such slow processes (Marković and Gros, 2014). A considerable body of research has focused upon the role of various forms of synaptic plasticity, including simple activity-dependent up- and down-regulation (de Arcangelis, 2008), activity- dependent synaptic plasticity (de Arcangelis et al., 2006), synaptic potentiation (Stepp et al., 2015), short-term synaptic depression through depletion of synaptic vesicles (Bonachela et al., 2010; Levina et al., 2014; Mihalas et al., 2014; Millman et al., 2010), Hebbian (Van Kessenich et al., 2016) and anti-Hebbian synaptic plasticity (Cowan et al., 2014; Magnasco et al., 2009), and spike-time dependent plasticity (de Andrade Costa et al., 2015; Rubinov et al., 2011). As with physical systems, the (relatively) slow synaptic plasticity serves to broaden the critical point to a broad, stable region. Other neurobiological processes have also been proposed, including balanced excitation-inhibition and network topology (Rubinov et al., 2011) and dynamic neuronal gain (Brochini et al., 2016). More recently, the role of energy build-up and dissipation in physical systems has been recast in critical neural systems as the replenishment and depletion of intracellular metabolic resources including Adenosine Triphosphate (ATP); (Roberts et al., 2014b; Stramaglia et al., 2015; Virkar et al., 2016). 4. Challenges and pitfalls of the criticality hypothesis Despite this recent emergence of criticality research in neuroscience, lessons learned in other branches of science raise important pitfalls and caveats. First, inferring the presence of scale-free statistics in neuroscience data has classically rested upon fitting a power-law (or Pareto) regression to the probability distribution of the size of the temporal or spatial fluctuations (Figure 1 and Figure 2). The statistical principles underlying this exercise were critiqued in a highly influential survey by Clauset and colleagues (Clauset et al., 2009). While a linear regression in double logarithmic coordinates can yield a fit that looks impressive, such a process is insensitive to the distribution of data at the right- hand tail of the distribution – the very region where the presence of a heavy-tailed power law needs to be rigorously tested. This is because the number of empirically measured samples found in the tail of the distribution is often too limited for robust inference. Another concern is that the samples of a cumulative distribution function are not independent, whereas regression assumes data independence. Clauset and colleagues developed a more principled approach based on maximum likelihood estimation to test and compare different statistical models of the data, including the power law, but also other candidate heavy tailed distributions including the log normal and stretched exponential forms (Clauset et al., 2009; Vuong, 1989). In brief, once the best fitting power-law parameters for an empirical dataset have been determined with maximum likelihood estimation, the goodness-of-fit between the fitted power law and empirical distribution is tested with the Kolmogorov-Smirnov (K-S) statistic. To this end, Clauset and colleagues proposed to randomly sample data from the fitted power law, independently fit a new power law to each of these new data samples and then evaluate the goodness-of-fit between the new samples and the new power laws. This is repeated many times to generate an empirical distribution of K-S statistics, which can then be used to compute a p-value for the K-S statistic corresponding to the observed data. If this p-value is significant, the randomly sampled data is a better fit to the power law than the observed data, and thus a power law should be excluded as an appropriate model. Otherwise, if the p-value is not significant, power-law behaviour is supported and the final step is to exclude other distributions as providing better evidence. Relative fits are typically computed between candidate distributions (lognormal, stretched exponentials, etc.) using log-likelihoods. Using this approach, Clauset et al. (2009) revisited several physical phenomena thought to have scale- free statistics and showed that several of these data were better explained by other long-tailed distributions, not power laws. Using these methods shows that the same holds true for many neuronal 19 Criticality in the brain Cocchi et al. 20 fluctuations (Roberts et al., 2014a). For example, it appears that fluctuating alpha rhythm follows a stretched exponential distribution, not a power law (Freyer et al., 2009). In turn, biophysical models suggest that the alpha rhythm arises from noise-driven multistability, rather than being generated by classic (super-) criticality (Freyer et al., 2011). A second caveat was issued by the empirical analyses and modelling work of Tomboul and Destexhe (Touboul and Destexhe, 2010, 2015) who showed that under certain situations, the aggregate behaviour of non-critical stochastic systems could yield irregular time series with power law statistics, albeit over a limited range. This is an important issue which also highlights the importance of null models for the different experimental methods (Farmer, 2015). These findings suggest that inferences regarding criticality based on the observation of power law scaling in empirical data should be made with caution, particularly if the scaling extends for less than two orders of magnitude or the slope of the power law is steep (Miller et al., 2009) - i.e. the scaling exponent  is greater than 2.5. A third caveat concerns other classes of interesting, emergent phenomena. Non-trivial, emergent dynamics can arise through other complex nonlinear phenomena. A classic example is that of so- called winnerless competition (Melbourne et al., 1989; Rabinovich et al., 2001) that has been proposed to underlie cognitive tasks such as animal gait (Golubitsky et al., 1999), perceptual rivalry (Ashwin and Lavric, 2010), and sequential decision-making (Rabinovich et al., 2008). Winnerless competition arises from metastable transitions along a sequence of unstable states (not unlike the drawings of M.C. Esher). Unlike criticality, the successive states are not weakly stable, but are unstable (Figure 4). In common with critical systems, noise plays a crucial role in a metastable system. However, the statistics of a metastable system are not power laws (Figure 4b). Rather the duration that the system dwells near each of its states varies in proportion to the logarithm of the noise amplitude (Ashwin et al., 2006). The temporal statistics therefore have a characteristic (and relatively short) time scale, and are not scale-free (Figure 4b). While multistable (Figure 4a) and metastable (Figure 4b) systems are therefore related (particularly in name!), their statistics are distinct and the underlying nonlinear causes differ. Unfortunately, the two terms are often used interchangeably. A number of final caveats pertain to the practical aspects of empirical data. First, the algorithm developed by Clauset et al. (2009) assumes that there is no upper bound to the empirical power-law distribution. This is a flawed assumption for most experimental data, which inevitably derive from a finite number of sensors, and may bias model selection. Recent work has revisited this assumption, developing methods that test the likelihood of a power law with a simple cut-off (Langlois et al., 2014; Shew et al., 2015). Second, as mentioned above, the range of many data tested for power law scaling often span less than two orders of magnitude, yielding data that is particularly sparse in the right-hand tail (precisely where power law scaling is most clearly expressed). Although the use of model estimation rather than linear regression partly mitigates this, disambiguating amongst the variety of candidate heavy-tailed distributions can only be reliably performed when the data scale over more than two orders of magnitude. Lengthy acquisitions may help here. For example, free- living activity patterns in humans derived from accelerometry recordings over seven consecutive days scale across four orders of magnitude: These allow for disambiguation of composite exponential and truncated power law distributions in active versus inactive periods of the day (Chapman et al., 2016). 20 Criticality in the brain Cocchi et al. 21 Figure 4: Different expressions of instability lead to different types of complex dynamics. (a) In a multistable system, noise drives a system erratically between different attractors. Because the system is briefly trapped in each basin of attraction, the time series shows a relatively long-tailed (stretched exponential) dwell distribution, here shown in linear-log coordinates (inset). (b) In a metastable system, there are no attractors, but rather a sequence of linked unstable fixed points. Because these are only weakly unstable, the system dwells in the neighbourhood of each, but does not show trapping. The sequential dwell times are therefore not long-tailed but show a characteristic time-scale corresponding to the peak in a gamma function shown here in linear coordinates (inset). In a critical system, a single fixed point is very weakly attracting or neutral. System noise leads to long and unstructured excursions corresponding to scale-free fluctuations and corresponding power-law statistics. Disambiguating these different underlying causes of complex dynamics 21 Criticality in the brain Cocchi et al. 22 the imperfect measurement process: In can be achieved with careful analyses of the system statistics, together with inversion of corresponding computational models. Finally, noisy fluctuations in physiological data do not only arise from the system of interest (i.e. neural activity), but also from the setting of neurophysiological data, these fluctuations consist of additive noise from physiological sources (such as muscular activity, cardiovascular influences) as well as artefacts due to extraneous effects (e.g. thermal scanner noise, head motion). In principle, such effects could lead to false positives in power- law evaluation. However, these inputs are generally uncorrelated and their summation therefore (according to the central limit theorem) likely to be Gaussian, not heavy-tailed. Nonetheless, care should be taken to disambiguate their contribution to any putative heavy-tailed system statistics in case one particular artefact (such as head movement in the scanner) dominates. Methods of doing this include: (i) taking independent measurements of these artefacts, such as taking empty scanner room recordings (Shriki et al., 2013) and ensuring that they do not possess the same statistics as attributed to the underlying neuronal system (Kitzbichler et al., 2009); (ii) using algorithms such as independent components analysis or source reconstruction to unmix neuronal fluctuations from physiological and measurement noise (Freyer et al., 2009); (iii) and using a formal inversion framework that formally accommodates measurements effects including spatial or temporal filtering and additive noise (Razi et al., 2015). These caveats highlight crucial points. While the application of criticality to neuroscience is an exciting field, progress needs to proceed with due caution. Analyses of neuroscience data for power laws first needs to consider other heavy-tailed candidate distributions. Second, inference should ultimately be based upon models of the causes of the observed statistics and avoid a direct inference of criticality that is based only upon data analysis. Third, computational models of neural systems that are based upon criticality should be tested closely against empirical data using appropriate frameworks (Daunizeau et al., 2009; Penny, 2012) that allow disambiguation against competing models that invoke other nonlinear mechanisms [for review, see Roberts et al. (2015)]. While these caveats highlight important limitations, experimental manipulations and recent theoretical developments offer new opportunities to explore the "criticality hypothesis". As noted at the outset of this review, the notion of universality (properties that transcend the details of a particular system and are thus found in many diverse settings) is one of the central appeals of criticality. Universal scaling laws - such as the presence of a scaling function that inter-relates the common underlying shape of critical fluctuations across temporal and spatial scales (Sethna et al., 2001; Zapperi et al., 2005) - can be extracted from data and subject to null hypothesis testing (Friedman et al., 2012; Roberts et al., 2014a). Relationships between the exponents of different (spatial and temporal) scaling laws, may also be derived from empirical data (Friedman et al., 2012; Roberts et al., 2014a) and benchmarked against the simple relationships predicted by the mathematical theory of phase transitions (Sethna et al., 2001). Also, as reviewed above, the theoretical advantages of criticality in neuronal models was demonstrated in a series of elegant empirical studies using pharmacological manipulation to sweep systems from subcritical to critical to supercritical (Gautam et al., 2015; Shew et al., 2011; Shew et al., 2009; Tagliazucchi et al., 2016): Showing a sudden change in a scaling law, corresponding to a peak in information capacity or dynamic range, provides convergent evidence for the occurrence of a phase transition in the underlying system. 22 Criticality in the brain Cocchi et al. 23 5. Emerging role of criticality in cognition Notwithstanding the aforementioned caveats, growing empirical and modelling research clearly supports the view that neural dynamics likely occur near critical instabilities. The recognition of the limitations of this new field simply shows that it has matured beyond the "proof of principle" stage (Feyerabend, 1993). The scene is thus set for the translation of criticality into cognitive and clinical brain research. In Section 3.1, we noted a canonical example of critical fluctuations near the transition from anti- syncopated to syncopated rhythmic finger tapping. Do critical dynamics generalize to other behaviours? Accelerometer-based analyses of free behaviour in humans (going about their everyday lives) shows that periods of inactivity exhibit power-law statistics (Nakamura et al., 2007), whose scaling coefficients differ between wake and sleep (Chapman et al., 2016) and differ again in major depression (Nakamura et al., 2008). Intriguingly, accumulating evidence suggests that periods of activity – although long-tailed – do not fit a power law, but rather a stretched exponential (e.g., inset Figure 4a, depicting the Weibull distribution). This shift in the statistical proprieties of a system as a function of context is emerging as a powerful tool to understand the neural principles supporting healthy and pathological brain functions. 5.1. Criticality in brain and behaviour In traditional cognitive neuroscience experiments, separate trials are typically treated as independent. However, they are not necessarily treated independently by research participants. Indeed it has been shown that reaction times between sequential trials show long-range correlations across a wide diversity of tasks (Palva and Palva, 2011; Thornton and Gilden, 2005). This is perhaps not surprising, given that, outside the scanner, human activity (Sreekumar et al., 2016; Sreekumar et al., 2014) and memory (Nielson et al., 2015) show a complex temporal structure that scales across many orders of magnitude. Detailed analyses of reaction time data in typical psychophysics experiments support the presence of a power law structure (Van Orden et al., 2005). However, definitive support for this position, as well as putative underlying causes such as criticality, do remain contentious (Heathcote et al., 2000; Heathcote et al., 1991; Wagenmakers et al., 2004). A rich literature on critical dynamics in brain and behaviour exists. However, with few notable exceptions (Jirsa et al., 1994; Kelso et al., 1992), these two streams of research have preceded largely in parallel. Several recent papers have stepped toward a unifying framework. By acquiring high density MEG/EEG data while participants performed an audiovisual threshold-stimulus detection task Palva et al. (Palva et al., 2013) showed that the critical exponents of scale-free neuronal dynamics correlate with the inter-individual variability in behavioural scaling laws. These correlations, show a specific anatomical pattern, with the combined delta and alpha frequency bands correlations mapping onto posterior parietal cortex. The same cortical areas appear involved when electrophysiological signals in the beta and gamma ranges are considered, with the addition of the cuneus and inferotemporal brain areas. Interestingly, a significant association between neural and behavioural long-range temporal correlations was also found in brain regions comprising the default mode brain network (Greicius et al., 2003). Notably, the brain-behaviour association found in task execution largely overlaps with an association between visual behavioural performances and scale- free dynamics observed in resting-state MEG data. This suggests that the temporal structure of endogenous neural dynamics are, to some extent, preserved during task performance. Similarly, the analysis of auditory task data showing that cortical activations organize as neural avalanches in both visual cognitive and resting-state contexts [(Arviv et al., 2015), see also (Shew et al., 2015)]. to be 23 Criticality in the brain Cocchi et al. 24 Moreover, near-critical neural dynamics may determine fluctuations in perceptual and cognitive processes. Overall, these findings highlight the strong interdependence between the near-critical neural processes characterizing resting-state and task-specific processes. 5.2. Suppression of criticality during task performance Using simultaneously acquired EEG and fMRI data, Fagerholm et al (Fagerholm et al., 2015) recently recapitulated previous findings of scale-free cortical dynamics at rest [e.g., (Palva et al., 2013)]. However, as participants engaged in tasks of increasing attentional load, the statistics of cortical activity appeared to shift increasingly further from the critical state. It was surmised that although criticality was important for the unconstrained "exploratory" resting-state, executing specific tasks required suppression of the associated variability under the influence of dorsal attentional networks (Fagerholm et al., 2015; Hellyer et al., 2014). In keeping with this notion, complementary computational analyses have suggested that altered anatomical connectivity, cognitive flexibility and information processing following traumatic brain injury are linked to a reduction in the variance of functional connectivity simulated by a model of coupled phase oscillators (Hellyer et al., 2014). The proposition that near-critical neural dynamics are central to the emergence of conscious cognition is consistent with the results of high-density electrocorticography recordings (ECoG) in primates, showing that loss of consciousness is characterized by a reduction in the number of eigenmodes (the number of "excited modes" in a dynamic system) that are close to instability (Solovey et al., 2015). Conversely, the return of consciousness appears to be accompanied by a corresponding increase in the number of eigenmodes close to instability. Recent detailed analyses of in vivo electrophysiological and two-photon recordings in rodents provide direct support for a clear association between the appearance of critical activity and the emergence of consciousness from anaesthesia (Bellay et al., 2015; Scott et al., 2014). Likewise, induction of unconsciousness through propofol anaeasthesia is accompanied by a loss of the signatures of criticality in fMRI data (Tagliazucchi et al., 2016). Non-invasive and invasive human neuroimaging work assessing neural critical states by the mean of neural avalanches analysis further support the link between near-critical neural regimes and behaviour (Palva et al., 2013; Priesemann et al., 2013; Tagliazucchi et al., 2012). For example, ECoG data collected from patients with refractory partial epilepsy were analysed as a function of vigilance (Priesemann et al., 2013). In general, the probability distribution frequency of neural avalanches followed a power law. However, slow wave sleep (SWS) was characterized by the most frequent occurrence of avalanches; unconstrained wakefulness showed an intermediate occurrence, while rapid eye movement sleep (REM) showed the fewest. These results are in line with previous findings from invasive recording in rats (Ribeiro et al., 2010) and highlight the close relationship between changes in near-critical dynamics and distinct mental states. It has been suggested that near-critical dynamics in the default mode operate to increase the "dynamic repertoire" of the brain when subjects are at rest (Deco and Jirsa, 2012; Deco et al., 2011) – that is, to increase the number of proximal "cognitive sets" or stored memories (Scarpetta and de Candia, 2013). This is consistent with modelling and empirical work, reviewed above, that reveals that criticality optimizes the entropy (Fagerholm et al., 2016) and information capacity (Shew et al., 2011) of spontaneous neuronal activity. Criticality at rest could also function to tune cortical dynamics to an optimal state of "expected uncertainty" concerning the nature of as yet unencountered sensory inputs (Friston et al., 2012b). That is, by facilitating a flexible, but also constrained degree of instability in sensory-perceptual systems, criticality could endow the cortex with the ability to adapt to volatile and non-stationary environmental fluctuations: Away from the critical point, internal fluctuations are 24 Criticality in the brain Cocchi et al. 25 stable and strongly damped, hence shrinking toward a baseline (mean) value. Close to the critical point, such fluctuations encompass a broad spectrum of scales, similar to the scale-invariant statistics of natural scenes (Field, 1987; Ruderman and Bialek, 1994). At a very fundamental level, this matching of internally-generated fluctuations to the statistics of external scenes, via criticality, revisits the notion that spontaneous cortical activity encodes an optimal model of the environment (Berkes et al., 2011) and uses this to deploy adaptive behavioural strategies (Fiser et al., 2010), including saccadic (Friston et al., 2012a) and heavy-tailed fixational eye movements (Roberts et al., 2013). Upon task execution - as uncertainty decreases - this unstable dynamic landscape could accordingly be suppressed, allowing stabilization of a single task-related attractor. Such proposals are consistent with the broader recognition that "brain noise" plays an adaptive role in health (Garrett et al., 2011) and ageing (McIntosh et al., 2010), but decreases during task execution (Churchland et al., 2010; Ponce-Alvarez et al., 2015). By analysing the statistics and spatiotemporal scaling laws of neural avalanches from ex vivo local field potential (LFP) recordings of visual cortex, Shew et al. (2015) showed that strong visual stimulation initially engendered super-critical dynamics. However, with the continuation of the visual input, these were quickly tuned to a critical state through neuronal adaptation. A simple neuronal model with the adaptation to input mediated by short-term synaptic depression was able to capture the switch from non-critical to critical activity with continuous input (Shew et al., 2015). These results highlight the importance of synaptic plasticity in switching between critical and non-critical regimes in order to facilitate perception and cognition. 6. Criticality in disease 6.1. Bifurcations and seizures Whereas the role of criticality in cognition is relatively nascent, casting seizures as dynamic disorders that arise out of critical instabilities is supported by an appreciable body of evidence (Da Silva et al., 2003; Meisel et al., 2012). The primary generalized seizures of childhood - Absence seizures – correspond to the presence of high amplitude 3 Hz spike-and-wave oscillations that appear and terminate equally quickly. These seizures have been modelled as critical bifurcations in corticothalamic loops by several groups (Destexhe and Sejnowski, 2001) although there is yet a lack of consensus as to whether these bifurcations are subcritical (Fröhlich et al., 2010; Suffczynski et al., 2004) or supercritical (Breakspear et al., 2006; Robinson et al., 2002). Generalised tonic-clonic seizures, which are associated with the progression of various high amplitude waveforms, have been modelled as a subcritical bifurcation (Breakspear et al., 2006). Accordingly, patients with epilepsy "reside" in pathological multistable dynamic regimes, due to neurophysiological disturbances and are occasionally perturbed into seizure dynamics (Breakspear et al., 2006). Fast epileptic activity recorded invasively with cortical surface recordings have also been subject to detailed models using critical bifurcations (Bartolomei et al., 2001; Wendling et al., 2005). These are proposed to reflect aberrant interactions between excitatory and inhibitory cortical neurons due to GABA-ergic dysfunction (Wendling et al., 2002). A detailed account of the complex progression of low frequency (DC-like) shifts in the electrical baseline and the nested high frequency oscillations that superimpose on these has recently been advanced using bifurcations: This "epileptor" - a set of equations that express the corresponding nested bifurcations - represents a comprehensive integration of slow and fast time scales underlying criticality and bifurcations in seizure activity (Jirsa et al., 2017; Jirsa et al., 2014). 25 Criticality in the brain Cocchi et al. 26 6.2. Crackling noise and neonatal burst-suppression Burst suppression (BS) is a class of electrocortical activity that occurs in preterm neonates, full-term newborn infants with encephalopathy, and following propofol anaesthesia. BS is characterised by high amplitude, irregular bursts that erratically punctuate a flat EEG trace (Niedermeyer et al., 1999) (Figure 5). Analyses of BS in preterm and encephalopathic infants show that the bursts are characterized by long-tailed, scale-free properties that stretch across a remarkable six orders of magnitude (Roberts et al., 2014a). While individual bursts are highly irregular, binning them according to their size and then averaging all bursts within each bin yields very smooth shapes that vary only slowly across scales (Roberts et al., 2014a) (Figure 5). The relationship between these scale-specific shapes are described by a so-called scaling function that speaks to the nature of criticality in all systems where it arises (Baldassarri et al., 2003; Papanikolaou et al., 2011). Notably, the exponents that describe the scaling and shape of the bursts pre-empt clinical outcomes in these neonates (Iyer et al., 2015b; Iyer et al., 2014). These shapes have been modelled by a term that describes metabolic depletion and replenishment in these critically vulnerable newborn brains (Roberts et al., 2014a). Despite the success in modelling BS in babies using a criticality framework, bursts in propofol anaesthesia do not have scale-free properties (Ching et al., 2010). In contrast, the effect of this type of anaesthesia in brain activity has been modelled with alternative (non-critical) dynamical mechanisms (Bojak et al., 2015). Figure 5: Critical statistics in neonatal hypoxic ischemic encephalopathy. (a) Scalp EEG channels show characteristic bursting pattern of activity with bursts erratically punctuating a flat EEG trace. (b) The area under each of these bursts (a measure of the energy discharged) shows power-law scaling over 3 orders of magnitude, with an exponential truncation at the far right. (c) The relationship between the length and duration of bursts also follows a power law (linear in double logarithmic coordinates. (d) Bayesian model selection shows that a truncated power law is easily the best model for these data. Panels (a) adapted from (Iyer et al., 2015a); panels (b), (c) and (d) adapted from (Roberts et al., 2014a). 26 Criticality in the brain Cocchi et al. 27 Establishing the presence of critical instabilities in neurological disorders opens a number of novel therapeutic and diagnostic windows [for review, see Coombes and Terry (2012)]. For example, in concert with computational modelling, treating seizures as transitions across bifurcations permits tracking the trajectory of the seizure through bifurcation space (Freestone et al., 2013; Nevado- Holgado et al., 2012), with a longer-term objective of seizure control (Nelson et al., 2011). Intriguingly, such control systems could use micro-stimulation to exploit the increased dynamic responsiveness of a system close to criticality as an "early warning system" (Scheffer et al., 2009): To our knowledge, incorporating this approach to seizure prediction within the framework of criticality theory has not yet been achieved. Likewise, deriving measures of near-criticality from clinical, bed- side recordings of burst suppression in preterm and encephalopathic neonates could be used as prognostic markers of long-term outcome, hence opening novel therapeutic windows for early intervention (Roberts et al., 2017). 6.3. Criticality and neuropsychiatric disorders Criticality, bifurcations and phase transitions have provided increasingly nuanced understandings of several major neurological conditions. It is intriguing to consider a possible role in neuropsychiatric conditions such as psychotic and affective disorders, although these are likely to be more subtle in their deviation from the "optimum critical point". Cognitive disturbances in schizophrenia, such as working memory, have been captured in computational models by changing the influence of fluctuations in a multistable landscape (Loh et al., 2007). Changes in large-scale resting-state dynamics seen in this disorder – such as a decrease in the global brain signal – have also been modelled by altering the balance of noise and stability in macroscopic neuronal activity (Yang et al., 2014). Schizophrenia is associated with less suppression of activity in the default-mode network when subjects engage in external tasks (Harrison et al., 2007; Meyer-Lindenberg et al., 2001; Nejad et al., 2011; Whitfield-Gabrieli et al., 2009). Dynamic insights into this loss of efficiency could draw from the theory of criticality. In particular, if adaptive cortical activity reflects a switch from sub-critical to super-critical activity under the influence of attention and arousal, then it follows that the precision of this process could be diminished in schizophrenia. Default activity would then be inadequately suppressed during attention to external tasks, while super-critical activity would not be optimally portioned into distinct spatiotemporal patterns. While this is clearly speculative, the tools of criticality provide new methodological tools and innovative directions for research in this often-intractable disorder. Whereas schizophrenia is characterized by disorganization – suggesting too little stability – core manifestations of other psychiatric disorders suggest too much stability. The ruminations of depressive disorders and the maladaptive pre-occupations of obsessive-compulsive disorder may arise in a system residing too deeply in the subcritical zone, preventing adaptive switching from interospective to exteroceptive states. Analysis of resting-state fMRI from patients with melancholia supports this view, showing a disturbance in the key dynamic parameters that modulate stability (Hyett et al., 2015a). Notably, when patients with melancholia view emotionally salient material, frontal attentional networks increase system stability, contrary to their influence in healthy controls (Hyett et al., 2015b). As noted above, critical instability is needed for a system to respond to the stream of stimuli. Here, too little instability is proposed to underlie the lack of reactivity. Although these are preliminary findings, the potential for criticality to provide fresh insights into tough neuropsychiatric research problems is again evident. It is quite possible that psychiatric phenotypes do not only reflect a simple failure of critical brain dynamics, but also their replacement by different dynamic processes in compensation (Breakspear et al., 2015). 27 Criticality in the brain Cocchi et al. 28 7. Summary Evidence for the widespread occurrence of criticality in nature, and its corresponding computational advantages, has triggered the interest of scientists in many different fields. The list of advantages associated with criticality spans many systems and different measurable quantities (Assis and Copelli, 2008; Boedecker et al., 2012; Deco et al., 2013; Gollo et al., 2013; Haldeman and Beggs, 2005; Hidalgo et al., 2014; Kastner et al., 2015; Legenstein and Maass, 2007; Livi et al., 2016; Mosqueiro and Maia, 2013; Publio et al., 2012; Schrauwen et al., 2009; Shew and Plenz, 2013). Despite the field traditionally developing outside of neuroscience, many of the most exciting findings now focus on brain dynamics. This body of work suggests that specific functions of the central nervous system may exploit optimal properties observed at criticality, representing a general functional property of the brain (Chialvo, 2006). What are the mechanisms that establish criticality in the brain? The obvious lack of an externally tunable parameter highlights the importance of self-organized criticality (SOC). As we reviewed above, SOC arises when the interactions between the elements (neurons) of a system are endowed with plasticity, or when there is a slow build-up and a fast release of energy. Synaptic processes such as frequency adaptation or spike-time dependent plasticity are clear candidates for SOC at the neuronal level. Homeostatic mechanisms, such as the hypothalamus-pituitary axis, govern metabolic processes in the body, and reflect the slow time-scales of energy accumulation and utilization. In this vein, neuronal criticality may extend beyond the central nervous system to incorporate the autonomic nervous system and corresponding interoceptive feedback loops to the midbrain and insula. Although criticality can arise from a fairly simple memory process, it seems unlikely to be confined to any single mechanism in the brain, but may reflect multiple processes operating for different goals (computation, adaptation, homeostasis). Likewise, changes in the environment, such as a scarcity of food sources, which mandate an (allostatic) change in the balance of energy metabolism may also engage SOC processes through a change in the critical set point. These more complex feedback systems are not classically thought of as SOC, but rather homeostatic, self-regulatory systems of the brain and body. Future work is required to understand if these broader homeo- and allostatic processes do follow the principles of SOC. Conversely, pathological failures of adaptive criticality may reflect disturbances in any of the supporting mechanisms, such as when a failure of balanced inhibition leads to a seizure, or when interoceptive disturbances lead to fatigue and depression (Petzschner et al., 2017). Criticality may also arise in pathological settings, such as if metabolic disturbances lead to an abnormal build-up and release of energy that would not normally occur: Burst suppression in the hypoxic newborn may be an example of this scenario (Roberts et al., 2015). The study of the brain as a complex, dynamic network is now flourishing. Modelling the inherently unstable, multiscale neuronal dynamics that generate flexible cognitive states is crucial to this endeavour. Although the notion of criticality is an attractive candidate, much work remains to be done to clarify the relationship between cognition and criticality. The increasing sophistication of theoretical models and their hypothesis-driven application to empirical data has to be tempered against the caveats that have recently emerged. Research that incorporates these lessons will shed new light on the dynamic brain networks that underlie action, perception and cognition in health and disease. 28 Criticality in the brain Cocchi et al. 29 Unifying models of brain function give sense to the results generated by the boom of explorative studies assessing the association between cognitive functions and neural dynamics. Without such models, there is a risk that the accumulation of experimental evidence will not substantially impact upon our understanding of the neural underpinnings of human cognition in health and disease. Recent studies suggest that the transition from critical to super-critical regimes may represent a general principle underlying the emergence of goal-directed behaviour. Those transitions may be moderated by slow integrators that are the highly connected hubs of the brain. In general, the framework of criticality challenges earlier prevailing conceptualizations of normal and pathological cognitive functions as emerging from discrete regions or static networks of brain regions. The science of criticality provides a new armoury of analytic techniques for basic and translational neuroscientists. The field is now sufficiently mature that inferring the presence of scale-free statistics from time series data should derive from formal likelihood tests and comparisons to other heavy- tailed processes. In addition, showing that the spatial and/or temporal statistics of a system are scale- free is now only the first step in emerging "best practice". Inferring that the underlying system is critical should ideally rest upon showing that the best model that can account for those statistics is one in which the model itself is perched at the cusp of criticality. Criticality should not become a catch-all term for everything that is complex or variable. It is only one amongst many possible causes of complex dynamics and care should be taken to disambiguate amongst these. Funding: L.C., L.L.G, A.Z and M.B. were supported by the Australian National Health Medical Research Council (L.C. APP1099082, L.L.G. APP1110975, A.Z. APP1047648, M.B. APP1037196). This work was also supported by the Australian Research Council Centre of Excellence for Integrative Brain Function (M.B., ARC Centre Grant CE140100007). Conflicts of Interest: The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. Acknowledgments: Dedicated to the memory of Walter Freeman whose seminal early work inspired us and shaped the field. 29 Criticality in the brain Cocchi et al. 30 8. References Arneodo, A., Coullet, P., Tresser, C., 1980. Occurence of strange attractors in three-dimensional Volterra equations. Physics Letters A 79, 259-263. Arviv, O., Goldstein, A., Shriki, O., 2015. Near-Critical Dynamics in Stimulus-Evoked Activity of the Human Brain and Its Relation to Spontaneous Resting-State Activity. The Journal of Neuroscience 35, 13927-13942. Ashwin, P., Burylko, O., Maistrenko, Y., Popovych, O., 2006. Extreme sensitivity to detuning for globally coupled phase oscillators. Physical review letters 96, 054102. Ashwin, P., Lavric, A., 2010. A low-dimensional model of binocular rivalry using winnerless competition. Physica D: Nonlinear Phenomena 239, 529-536. Assis, V.R., Copelli, M., 2008. Dynamic range of hypercubic stochastic excitable media. Physical Review E 77, 011923. Bak, P., 1990. Self-organized criticality. Physica A: Statistical Mechanics and its Applications 163, 403-409. Bak, P., Tang, C., Wiesenfeld, K., 1987. Self-organized criticality: An explanation of the 1/f noise. Physical review letters 59, 381-384. Bak, P., Tang, C., Wiesenfeld, K., 1988. Self-Organized Criticality. Phys Rev A 38, 364-374. Baldassarri, A., Colaiori, F., Castellano, C., 2003. Average shape of a fluctuation: Universality in excursions of stochastic processes. Physical review letters 90, 060601. Bartolomei, F., Wendling, F., Bellanger, J.-J., Régis, J., Chauvel, P., 2001. Neural networks involving the medial temporal structures in temporal lobe epilepsy. Clinical neurophysiology 112, 1746- 1760. Bassett, D.S., Wymbs, N.F., Rombach, M.P., Porter, M.A., Mucha, P.J., Grafton, S.T., 2013. Task- based core-periphery organization of human brain dynamics. PLoS computational biology 9, e1003171. Beggs, J.M., 2007. Neuronal avalanche. Scholarpedia 2, 1344. Beggs, J.M., Plenz, D., 2003. Neuronal avalanches in neocortical circuits. The Journal of neuroscience : the official journal of the Society for Neuroscience 23, 11167-11177. Beggs, J.M., Timme, N., 2012. Being critical of criticality in the brain. Frontiers in physiology 3, 163. Bellay, T., Klaus, A., Seshadri, S., Plenz, D., 2015. Irregular spiking of pyramidal neurons organizes as scale-invariant neuronal avalanches in the awake state. Elife 4, e07224. Berkes, P., Orbán, G., Lengyel, M., Fiser, J., 2011. Spontaneous cortical activity reveals hallmarks of an optimal internal model of the environment. Science 331, 83-87. Berryman, A.A., 1992. The Orgins and Evolution of Predator‐Prey Theory. Ecology 73, 1530-1535. Bertschinger, N., Natschläger, T., 2004. Real-time computation at the edge of chaos in recurrent neural networks. Neural computation 16, 1413-1436. Biswal, B.B., Mennes, M., Zuo, X.N., Gohel, S., Kelly, C., Smith, S.M., Beckmann, C.F., Adelstein, J.S., Buckner, R.L., Colcombe, S., Dogonowski, A.M., Ernst, M., Fair, D., Hampson, M., Hoptman, M.J., Hyde, J.S., Kiviniemi, V.J., Kotter, R., Li, S.J., Lin, C.P., Lowe, M.J., Mackay, C., Madden, D.J., Madsen, K.H., Margulies, D.S., Mayberg, H.S., McMahon, K., Monk, C.S., Mostofsky, S.H., Nagel, B.J., Pekar, J.J., Peltier, S.J., Petersen, S.E., Riedl, V., Rombouts, S.A., Rypma, B., Schlaggar, B.L., Schmidt, S., Seidler, R.D., Siegle, G.J., Sorg, C., Teng, G.J., Veijola, J., Villringer, A., Walter, M., Wang, L., Weng, X.C., Whitfield- Gabrieli, S., Williamson, P., Windischberger, C., Zang, Y.F., Zhang, H.Y., Castellanos, F.X., Milham, M.P., 2010. Toward discovery science of human brain function. Proceedings of the National Academy of Sciences of the United States of America 107, 4734-4739. Boedecker, J., Obst, O., Lizier, J.T., Mayer, N.M., Asada, M., 2012. Information processing in echo state networks at the edge of chaos. Theory in Biosciences 131, 205-213. Bojak, I., Stoyanov, Z.V., Liley, D.T., 2015. Emergence of spatially heterogeneous burst suppression in a neural field model of electrocortical activity. Frontiers in systems neuroscience 9. Bonachela, J.A., De Franciscis, S., Torres, J.J., Munoz, M.A., 2010. Self-organization without conservation: are neuronal avalanches generically critical? Journal of Statistical Mechanics: Theory and Experiment 2010, P02015. 30 Criticality in the brain Cocchi et al. 31 Breakspear, M., 2001. Perception of odors by a nonlinear model of the olfactory bulb. International Journal of Neural Systems 11, 101-124. Breakspear, M., Roberts, G., Green, M.J., Nguyen, V.T., Frankland, A., Levy, F., Lenroot, R., Mitchell, P.B., 2015. Network dysfunction of emotional and cognitive processes in those at genetic risk of bipolar disorder. Brain, awv261. Breakspear, M., Roberts, J., Terry, J.R., Rodrigues, S., Mahant, N., Robinson, P., 2006. A unifying explanation of primary generalized seizures through nonlinear brain modeling and bifurcation analysis. Cerebral Cortex 16, 1296-1313. Bressler, S.L., Kelso, J.S., 2001. Cortical coordination dynamics and cognition. Trends in Cognitive Sciences 5, 26-36. Brochini, L., de Andrade Costa, A., Abadi, M., Roque, A.C., Stolfi, J., Kinouchi, O., 2016. Phase transitions and self-organized criticality in networks of stochastic spiking neurons. Scientific reports 6. Burridge, R., Knopoff, L., 1967. Model and theoretical seismicity. Bull. Seismol. Soc. Am. 57, 3411– 3471. Camalet, S., Duke, T., Jülicher, F., Prost, J., 2000. Auditory sensitivity provided by self-tuned critical oscillations of hair cells. Proceedings of the National Academy of Sciences 97, 3183-3188. Carlson, J., Langer, J., 1989. Mechanical model of an earthquake fault. Phys Rev A 40, 6470. Cavagna, A., Cimarelli, A., Giardina, I., Parisi, G., Santagati, R., Stefanini, F., Viale, M., 2010. Scale- free correlations in starling flocks. Proceedings of the National Academy of Sciences 107, 11865-11870. Chapman, J.J., Roberts, J.A., Nguyen, V.T., Breakspear, M., 2016. Quantification of free-living activity patterns using accelerometry in adults with mental illness. arXiv preprint arXiv:1612.04471. Chialvo, D.R., 2006. The brain near the edge. arXiv preprint q-bio/0610041. Chialvo, D.R., 2010. Emergent complex neural dynamics. Nature Physics 6, 744-750. Chialvo, D.R., 2012. Critical brain dynamics at large scale. arXiv preprint arXiv:1210.3632. Ching, S., Cimenser, A., Purdon, P.L., Brown, E.N., Kopell, N.J., 2010. Thalamocortical model for a propofol-induced α-rhythm associated with loss of consciousness. Proceedings of the National Academy of Sciences 107, 22665-22670. Churchland, M.M., Yu, B.M., Cunningham, J.P., Sugrue, L.P., Cohen, M.R., Corrado, G.S., Newsome, W.T., Clark, A.M., Hosseini, P., Scott, B.B., Bradley, D.C., Smith, M.A., Kohn, A., Movshon, J.A., Armstrong, K.M., Moore, T., Chang, S.W., Snyder, L.H., Lisberger, S.G., Priebe, N.J., Finn, I.M., Ferster, D., Ryu, S.I., Santhanam, G., Sahani, M., Shenoy, K.V., 2010. Stimulus onset quenches neural variability: a widespread cortical phenomenon. Nature neuroscience 13, 369-378. Clauset, A., Shalizi, C.R., Newman, M.E., 2009. Power-law distributions in empirical data. SIAM review 51, 661-703. Cocchi, L., Sale, M.V., L, L.G., Bell, P.T., Nguyen, V.T., Zalesky, A., Breakspear, M., Mattingley, J.B., 2016. A hierarchy of timescales explains distinct effects of local inhibition of primary visual cortex and frontal eye fields. Elife 5. Cocchi, L., Zalesky, A., Fornito, A., Mattingley, J.B., 2013. Dynamic cooperation and competition between brain systems during cognitive control. Trends Cogn Sci 17, 493-501. Coombes, S., Terry, J.R., 2012. The dynamics of neurological disease: integrating computational, experimental and clinical neuroscience. European Journal of Neuroscience 36, 2118-2120. Copelli, M., 2014. Criticality at Work: How Do Critical Networks Respond to Stimuli? Criticality in Neural Systems, 347-364. Cote, P., Meisel, L., 1991. Self-organized criticality and the Barkhausen effect. Physical review letters 67, 1334. Cowan, J.D., Neuman, J., van Drongelen, W., 2014. Self-organized criticality and near criticality in neural networks. Criticality in Neural Systems, 465-484. Crutchfield, J.P., Young, K. (1988) Computation at the onset of chaos. In: The Santa Fe Institute, Westview. Citeseer. 31 Criticality in the brain Cocchi et al. 32 Da Silva, F.L., Blanes, W., Kalitzin, S.N., Parra, J., Suffczynski, P., Velis, D.N., 2003. Epilepsies as dynamical diseases of brain systems: basic models of the transition between normal and epileptic activity. Epilepsia 44, 72-83. Damoiseaux, J.S., Rombouts, S.A., Barkhof, F., Scheltens, P., Stam, C.J., Smith, S.M., Beckmann, C.F., 2006. Consistent resting-state networks across healthy subjects. Proceedings of the National Academy of Sciences of the United States of America 103, 13848-13853. Daunizeau, J., Friston, K.J., Kiebel, S.J., 2009. Variational Bayesian identification and prediction of stochastic nonlinear dynamic causal models. Physica D: Nonlinear Phenomena 238, 2089- 2118. de Andrade Costa, A., Copelli, M., Kinouchi, O., 2015. Can dynamical synapses produce true self- organized criticality? Journal of Statistical Mechanics: Theory and Experiment 2015, P06004. de Arcangelis, L. 2008. Activity-Dependent Model for Neuronal Avalanches. In: Aspects of Physical Biology. pp. 215-230. Springer. de Arcangelis, L., Perrone-Capano, C., Herrmann, H.J., 2006. Self-organized criticality model for brain plasticity. Physical review letters 96, 028107. De Carvalho, J.X., Prado, C.P., 2000. Self-organized criticality in the Olami-Feder-Christensen model. Physical review letters 84, 4006. Deco, G., Jirsa, V., McIntosh, A.R., Sporns, O., Kotter, R., 2009. Key role of coupling, delay, and noise in resting brain fluctuations. Proceedings of the National Academy of Sciences of the United States of America 106, 10302-10307. Deco, G., Jirsa, V.K., 2012. Ongoing cortical activity at rest: criticality, multistability, and ghost attractors. The Journal of neuroscience 32, 3366-3375. Deco, G., Jirsa, V.K., McIntosh, A.R., 2011. Emerging concepts for the dynamical organization of resting-state activity in the brain. Nature Reviews Neuroscience 12, 43-56. Deco, G., Jirsa, V.K., McIntosh, A.R., 2013. Resting brains never rest: computational insights into potential cognitive architectures. Trends in neurosciences 36, 268-274. Destexhe, A., Sejnowski, T.J., 2001. Thalamocortical Assemblies: How Ion Channels, Single Neurons and Large-Scale Networks Organize Sleep Oscillations. Eguiluz, V.M., Chialvo, D.R., Cecchi, G.A., Baliki, M., Apkarian, A.V., 2005. Scale-free brain functional networks. Physical review letters 94. Eguíluz, V.M., Ospeck, M., Choe, Y., Hudspeth, A., Magnasco, M.O., 2000. Essential nonlinearities in hearing. Physical review letters 84, 5232. Fagerholm, E.D., Lorenz, R., Scott, G., Dinov, M., Hellyer, P.J., Mirzaei, N., Leeson, C., Carmichael, D.W., Sharp, D.J., Shew, W.L., Leech, R., 2015. Cascades and cognitive state: focused attention incurs subcritical dynamics. The Journal of neuroscience : the official journal of the Society for Neuroscience 35, 4626-4634. Fagerholm, E.D., Scott, G., Shew, W.L., Song, C., Leech, R., Knöpfel, T., Sharp, D.J., 2016. Cortical Entropy, Mutual Information and Scale-Free Dynamics in Waking Mice. Cerebral Cortex. Farmer, S., 2015. Comment on "Broadband Criticality of Human Brain Network Synchronization" by Kitzbichler MG, Smith ML, Christensen SR, Bullmore E (2009) PLoS Comput Biol 5: e1000314. Feyerabend, P., 1993. Against method. Verso. Field, D.J., 1987. Relations between the statistics of natural images and the response properties of cortical cells. JOSA A 4, 2379-2394. Fiser, J., Berkes, P., Orbán, G., Lengyel, M., 2010. Statistically optimal perception and learning: from behavior to neural representations. Trends in cognitive sciences 14, 119-130. Fornito, A., Zalesky, A., Breakspear, M., 2015. The connectomics of brain disorders. Nature reviews. Neuroscience 16, 159-172. Fox, M.D., Snyder, A.Z., Vincent, J.L., Corbetta, M., Van Essen, D.C., Raichle, M.E., 2005. The human brain is intrinsically organized into dynamic, anticorrelated functional networks. Proceedings of the National Academy of Sciences of the United States of America 102, 9673- 9678. Freeman, W.J., 1987. Simulation of chaotic EEG patterns with a dynamic model of the olfactory system. Biological cybernetics 56, 139-150. 32 Criticality in the brain Cocchi et al. 33 Freeman, W.J., 1991. The physiology of perception. Scientific American 264, 78-85. Freestone, D.R., Long, S.N., Frey, S., Stypulkowski, P.H., Giftakis, J.E., Cook, M.J. (2013) A method for actively tracking excitability of brain networks using a fully implantable monitoring system. In: Engineering in Medicine and Biology Society (EMBC), 2013 35th Annual International Conference of the IEEE. pp. 6151-6154. IEEE. Freyer, F., Aquino, K., Robinson, P.A., Ritter, P., Breakspear, M., 2009. Bistability and non-Gaussian fluctuations in spontaneous cortical activity. The Journal of neuroscience : the official journal of the Society for Neuroscience 29, 8512-8524. Freyer, F., Roberts, J.A., Becker, R., Robinson, P.A., Ritter, P., Breakspear, M., 2011. Biophysical mechanisms of multistability in resting-state cortical rhythms. The Journal of neuroscience : the official journal of the Society for Neuroscience 31, 6353-6361. Freyer, F., Roberts, J.A., Ritter, P., Breakspear, M., 2012. A canonical model of multistability and scale-invariance in biological systems. PLoS computational biology 8, e1002634. Friedman, N., Ito, S., Brinkman, B.A., Shimono, M., DeVille, R.L., Dahmen, K.A., Beggs, J.M., Butler, T.C., 2012. Universal critical dynamics in high resolution neuronal avalanche data. Physical review letters 108, 208102. Friston, K., Adams, R., Perrinet, L., Breakspear, M., 2012a. Perceptions as hypotheses: saccades as experiments. Frontiers in psychology 3, 151. Friston, K., Breakspear, M., Deco, G., 2012b. Perception and self-organized instability. Front Comput Neurosci 6. Friston, K.J., 2000. The labile brain. II. Transients, complexity and selection. Philosophical Transactions of the Royal Society of London B: Biological Sciences 355, 237-252. Fröhlich, F., Sejnowski, T.J., Bazhenov, M., 2010. Network bistability mediates spontaneous transitions between normal and pathological brain states. The Journal of neuroscience 30, 10734-10743. Gal, A., Marom, S., 2013. Self-organized criticality in single-neuron excitability. Physical Review E 88. Garrett, D.D., Kovacevic, N., McIntosh, A.R., Grady, C.L., 2011. The importance of being variable. The Journal of neuroscience 31, 4496-4503. Gatlin, L.L., 1972. Information theory and the living system. Gautam, S.H., Hoang, T.T., McClanahan, K., Grady, S.K., Shew, W.L., 2015. Maximizing sensory dynamic range by tuning the cortical state to criticality. PLoS computational biology 11, e1004576. Gireesh, E.D., Plenz, D., 2008. Neuronal avalanches organize as nested theta-and beta/gamma- oscillations during development of cortical layer 2/3. Proceedings of the National Academy of Sciences 105, 7576-7581. Gollo, L.L., 2017. Coexistence of critical sensitivity and subcritical specificity can yield optimal population coding. arXiv preprint arXiv:1707.04484. Gollo, L.L., Breakspear, M., 2014. The frustrated brain: from dynamics on motifs to communities and networks. Philosophical transactions of the Royal Society of London. Series B, Biological sciences 369. Gollo, L.L., Copelli, M., Roberts, J.A., 2016. Diversity improves performance in excitable networks. PeerJ 4, e1912. Gollo, L.L., Kinouchi, O., Copelli, M., 2013. Single-neuron criticality optimizes analog dendritic computation. Scientific reports 3. Gollo, L.L., Mirasso, C., Eguíluz, V.M., 2012. Signal integration enhances the dynamic range in neuronal systems. Physical Review E 85, 040902. Gollo, L.L., Roberts, J.A., Cocchi, L., 2017. Mapping how local perturbations influence systems-level brain dynamics. Neuroimage. Gollo, L.L., Zalesky, A., Hutchison, R.M., van den Heuvel, M., Breakspear, M., 2015. Dwelling quietly in the rich club: brain network determinants of slow cortical fluctuations. Philosophical transactions of the Royal Society of London. Series B, Biological sciences 370. Golubitsky, M., Stewart, I., Buono, P.L., Collins, J.J., 1999. Symmetry in locomotor central pattern generators and animal gaits. Nature 401, 693-695. 33 Criticality in the brain Cocchi et al. 34 Greenfield, E., Lecar, H., 2001. Mutual information in a dilute, asymmetric neural network model. Physical Review E 63, 041905. Greicius, M.D., Krasnow, B., Reiss, A.L., Menon, V., 2003. Functional connectivity in the resting brain: a network analysis of the default mode hypothesis. Proceedings of the National Academy of Sciences of the United States of America 100, 253-258. Haimovici, A., Tagliazucchi, E., Balenzuela, P., Chialvo, D.R., 2013. Brain organization into resting state networks emerges at criticality on a model of the human connectome. Physical review letters 110, 178101. Haken, H., Kelso, J.S., Bunz, H., 1985. A theoretical model of phase transitions in human hand movements. Biological cybernetics 51, 347-356. Haldeman, C., Beggs, J.M., 2005. Critical branching captures activity in living neural networks and maximizes the number of metastable states. Physical review letters 94, 058101. Harrison, B.J., Yucel, M., Pujol, J., Pantelis, C., 2007. Task-induced deactivation of midline cortical regions in schizophrenia assessed with fMRI. Schizophrenia research 91, 82-86. Hasson, U., Chen, J., Honey, C.J., 2015. Hierarchical process memory: memory as an integral component of information processing. Trends Cogn Sci 19, 304-313. Hasson, U., Yang, E., Vallines, I., Heeger, D.J., Rubin, N., 2008. A hierarchy of temporal receptive windows in human cortex. The Journal of neuroscience : the official journal of the Society for Neuroscience 28, 2539-2550. He, B.J., 2013. Scale-free dynamics and critical phenomena in cortical activity. Frontiers E-books. He, B.J., Zempel, J.M., Snyder, A.Z., Raichle, M.E., 2010. The temporal structures and functional significance of scale-free brain activity. Neuron 66, 353-369. Hearne, L., Cocchi, L., Zalesky, A., Mattingley, J.B., 2015. Interactions between default mode and control networks as a function of increasing cognitive reasoning complexity. Human brain mapping. Heathcote, A., Brown, S., Mewhort, D., 2000. The power law repealed: The case for an exponential law of practice. Psychonomic bulletin & review 7, 185-207. Heathcote, A., Popiel, S.J., Mewhort, D., 1991. Analysis of response time distributions: An example using the Stroop task. Psychological Bulletin 109, 340. Hellyer, P.J., Shanahan, M., Scott, G., Wise, R.J., Sharp, D.J., Leech, R., 2014. The control of global brain dynamics: opposing actions of frontoparietal control and default mode networks on attention. The Journal of neuroscience : the official journal of the Society for Neuroscience 34, 451-461. Hesse, J., Gross, T., 2015. Self-organized criticality as a fundamental property of neural systems. Criticality as a signature of healthy neural systems: multi-scale experimental and computational studies. Hidalgo, J., Grilli, J., Suweis, S., Muñoz, M.A., Banavar, J.R., Maritan, A., 2014. Information-based fitness and the emergence of criticality in living systems. Proceedings of the National Academy of Sciences 111, 10095-10100. Honey, C.J., Kotter, R., Breakspear, M., Sporns, O., 2007. Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences of the United States of America 104, 10240-10245. Honey, C.J., Thesen, T., Donner, T.H., Silbert, L.J., Carlson, C.E., Devinsky, O., Doyle, W.K., Rubin, N., Heeger, D.J., Hasson, U., 2012. Slow cortical dynamics and the accumulation of information over long timescales. Neuron 76, 423-434. Houle, P.A., Sethna, J.P., 1996. Acoustic emission from crumpling paper. Physical Review E 54, 278. Hudspeth, A., 2008. Making an effort to listen: mechanical amplification in the ear. Neuron 59, 530- 545. Hyett, M.P., Breakspear, M.J., Friston, K.J., Guo, C.C., Parker, G.B., 2015a. Disrupted effective connectivity of cortical systems supporting attention and interoception in melancholia. JAMA psychiatry 72, 350-358. Hyett, M.P., Parker, G.B., Guo, C.C., Zalesky, A., Nguyen, V.T., Yuen, T., Breakspear, M., 2015b. Scene unseen: Disrupted neuronal adaptation in melancholia during emotional film viewing. NeuroImage: Clinical 9, 660-667. 34 Criticality in the brain Cocchi et al. 35 Iyer, K.K., Roberts, J.A., Hellstrom-Westas, L., Wikstrom, S., Hansen Pupp, I., Ley, D., Breakspear, M., Vanhatalo, S., 2015a. Early Detection of Preterm Intraventricular Hemorrhage From Clinical Electroencephalography. Crit Care Med 43, 2219-2227. Iyer, K.K., Roberts, J.A., Hellstrom-Westas, L., Wikstrom, S., Hansen Pupp, I., Ley, D., Vanhatalo, S., Breakspear, M., 2015b. Cortical burst dynamics predict clinical outcome early in extremely preterm infants. Brain 138, 2206-2218. Iyer, K.K., Roberts, J.A., Metsaranta, M., Finnigan, S., Breakspear, M., Vanhatalo, S., 2014. Novel features of early burst suppression predict outcome after birth asphyxia. Annals of clinical and translational neurology 1, 209-214. Jirsa, V.K., Friedrich, R., Haken, H., Kelso, J.S., 1994. A theoretical model of phase transitions in the human brain. Biological cybernetics 71, 27-35. Jirsa, V.K., Proix, T., Perdikis, D., Woodman, M.M., Wang, H., Gonzalez-Martinez, J., Bernard, C., Bénar, C., Guye, M., Chauvel, P., 2017. The virtual epileptic patient: individualized whole- brain models of epilepsy spread. Neuroimage 145, 377-388. Jirsa, V.K., Stacey, W.C., Quilichini, P.P., Ivanov, A.I., Bernard, C., 2014. On the nature of seizure dynamics. Brain 137, 2210-2230. Kastner, D.B., Baccus, S.A., Sharpee, T.O., 2015. Critical and maximally informative encoding between neural populations in the retina. Proceedings of the National Academy of Sciences 112, 2533-2538. Kauffman, S.A., Johnsen, S., 1991. Coevolution to the edge of chaos: Coupled fitness landscapes, poised states, and coevolutionary avalanches. Journal of theoretical biology 149, 467-505. Kello, C.T., Brown, G.D., Ferrer-i-Cancho, R., Holden, J.G., Linkenkaer-Hansen, K., Rhodes, T., Van Orden, G.C., 2010. Scaling laws in cognitive sciences. Trends in cognitive sciences 14, 223- 232. Kelso, J., 1984. Phase transitions and critical behavior in human bimanual coordination. American Journal of Physiology-Regulatory, Integrative and Comparative Physiology 246, R1000- R1004. Kelso, J., 2014. The dynamic brain in action: Coordinative structures, criticality, and coordination dynamics. Criticality in Neural Systems, 67-104. Kelso, J., Bressler, S., Buchanan, S., DeGuzman, G., Ding, M., Fuchs, A., Holroyd, T., 1992. A phase transition in human brain and behavior. Physics Letters A 169, 134-144. Kelso, J., Scholz, J., Schöner, G., 1986. Nonequilibrium phase transitions in coordinated biological motion: critical fluctuations. Physics Letters A 118, 279-284. Kelso, J.S., Clark, J.E., 1982. The development of movement control and coordination. John Wiley & Sons. Kim, S., Park, S.H., Ryu, C.S., 1997. Noise-enhanced multistability in coupled oscillator systems. Physical review letters 78, 1616. Kinouchi, O., Copelli, M., 2006. Optimal dynamical range of excitable networks at criticality. Nature physics 2, 348-351. Kitzbichler, M.G., Smith, M.L., Christensen, S.R., Bullmore, E., 2009. Broadband criticality of human brain network synchronization. PLoS computational biology 5, e1000314. Kosterlitz, J.M., Thouless, D.J., 1973. Ordering, metastability and phase transitions in two- dimensional systems. Journal of Physics C: Solid State Physics 6, 1181. Kramer, E.M., Lobkovsky, A.E., 1996. Universal power law in the noise from a crumpled elastic sheet. Physical Review E 53, 1465. Langlois, D., Cousineau, D., Thivierge, J.-P., 2014. Maximum likelihood estimators for truncated and censored power-law distributions show how neuronal avalanches may be misevaluated. Physical Review E 89, 012709. Langton, C.G., 1990. Computation at the edge of chaos: phase transitions and emergent computation. Physica D: Nonlinear Phenomena 42, 12-37. Larremore, D.B., Shew, W.L., Ott, E., Sorrentino, F., Restrepo, J.G., 2014. Inhibition causes ceaseless dynamics in networks of excitable nodes. Physical review letters 112, 138103. Larremore, D.B., Shew, W.L., Restrepo, J.G., 2011. Predicting criticality and dynamic range in complex networks: effects of topology. Physical review letters 106, 058101. 35 Criticality in the brain Cocchi et al. 36 Leech, R., Braga, R., Sharp, D.J., 2012. Echoes of the brain within the posterior cingulate cortex. The Journal of neuroscience : the official journal of the Society for Neuroscience 32, 215-222. Legenstein, R., Maass, W., 2007. Edge of chaos and prediction of computational performance for neural circuit models. Neural Networks 20, 323-334. Levina, A., Herrmann, J.M., Geisel, T., 2014. Theoretical Neuroscience of Self‐Organized Criticality: From Formal Approaches to Realistic Models. Criticality in Neural Systems, 417-436. Linkenkaer-Hansen, K., Nikouline, V.V., Palva, J.M., Ilmoniemi, R.J., 2001. Long-range temporal correlations and scaling behavior in human brain oscillations. The Journal of neuroscience : the official journal of the Society for Neuroscience 21, 1370-1377. Livi, L., Bianchi, F.M., Alippi, C., 2016. Determination of the edge of criticality in echo state networks through Fisher information maximization. arXiv preprint arXiv:1603.03685. Loh, M., Rolls, E.T., Deco, G., 2007. A dynamical systems hypothesis of schizophrenia. Lu, E.T., Hamilton, R.J., McTiernan, J., Bromund, K.R., 1993. Solar flares and avalanches in driven dissipative systems. The Astrophysical Journal 412, 841-852. Magnasco, M.O., Piro, O., Cecchi, G.A., 2009. Self-tuned critical anti-Hebbian networks. Physical review letters 102, 258102. Maoiléidigh, D.Ó., Nicola, E.M., Hudspeth, A., 2012. The diverse effects of mechanical loading on active hair bundles. Proceedings of the National Academy of Sciences 109, 1943-1948. Marković, D., Gros, C., 2014. Power laws and self-organized criticality in theory and nature. Physics Reports 536, 41-74. Markram, H., Muller, E., Ramaswamy, S., Reimann, M.W., Abdellah, M., Sanchez, C.A., Ailamaki, A., Alonso-Nanclares, L., Antille, N., Arsever, S., Kahou, G.A., Berger, T.K., Bilgili, A., Buncic, N., Chalimourda, A., Chindemi, G., Courcol, J.D., Delalondre, F., Delattre, V., Druckmann, S., Dumusc, R., Dynes, J., Eilemann, S., Gal, E., Gevaert, M.E., Ghobril, J.P., Gidon, A., Graham, J.W., Gupta, A., Haenel, V., Hay, E., Heinis, T., Hernando, J.B., Hines, M., Kanari, L., Keller, D., Kenyon, J., Khazen, G., Kim, Y., King, J.G., Kisvarday, Z., Kumbhar, P., Lasserre, S., Le Be, J.V., Magalhaes, B.R., Merchan-Perez, A., Meystre, J., Morrice, B.R., Muller, J., Munoz-Cespedes, A., Muralidhar, S., Muthurasa, K., Nachbaur, D., Newton, T.H., Nolte, M., Ovcharenko, A., Palacios, J., Pastor, L., Perin, R., Ranjan, R., Riachi, I., Rodriguez, J.R., Riquelme, J.L., Rossert, C., Sfyrakis, K., Shi, Y., Shillcock, J.C., Silberberg, G., Silva, R., Tauheed, F., Telefont, M., Toledo-Rodriguez, M., Trankler, T., Van Geit, W., Diaz, J.V., Walker, R., Wang, Y., Zaninetta, S.M., DeFelipe, J., Hill, S.L., Segev, I., Schurmann, F., 2015. Reconstruction and Simulation of Neocortical Microcircuitry. Cell 163, 456-492. McClure Jr, J.C., Schroder, K., 1976. The magnetic Barkhausen effect. Critical Reviews in Solid State and Material Sciences 6, 45-83. McIntosh, A.R., Kovacevic, N., Lippe, S., Garrett, D., Grady, C., Jirsa, V., 2010. The development of a noisy brain. Arch Ital Biol 148, 323-337. Meisel, C., Storch, A., Hallmeyer-Elgner, S., Bullmore, E., Gross, T., 2012. Failure of adaptive self- organized criticality during epileptic seizure attacks. PLoS computational biology 8, e1002312. Meisel, L., Cote, P., 1992. Power laws, flicker noise, and the Barkhausen effect. Physical Review B 46, 10822. Melbourne, I., Chossat, P., Golubitsky, M., 1989. Heteroclinic cycles involving periodic solutions in mode interactions with O (2) symmetry. Proceedings of the Royal Society of Edinburgh: Section A Mathematics 113, 315-345. Meyer-Lindenberg, A., Poline, J.B., Kohn, P.D., Holt, J.L., Egan, M.F., Weinberger, D.R., Berman, K.F., 2001. Evidence for abnormal cortical functional connectivity during working memory in schizophrenia. The American journal of psychiatry 158, 1809-1817. Mihalas, S., Millman, D., Iyer, R., Kirkwood, A., Niebur, E., 2014. Nonconservative Neuronal Networks During Up‐States Self‐Organize Near Critical Points. Criticality in Neural Systems, 437-464. Miller, K.J., Sorensen, L.B., Ojemann, J.G., den Nijs, M., 2009. Power-law scaling in the brain surface electric potential. PLoS computational biology 5, e1000609. 36 Criticality in the brain Cocchi et al. 37 Millman, D., Mihalas, S., Kirkwood, A., Niebur, E., 2010. Self-organized criticality occurs in non- conservative neuronal networks during/up/'states. Nature physics 6, 801-805. Misic, B., Betzel, R.F., Nematzadeh, A., Goni, J., Griffa, A., Hagmann, P., Flammini, A., Ahn, Y.Y., Sporns, O., 2015. Cooperative and Competitive Spreading Dynamics on the Human Connectome. Neuron 86, 1518-1529. Mitchell, M., Hraber, P., Crutchfield, J.P., 1993. Revisiting the edge of chaos: Evolving cellular automata to perform computations. arXiv preprint adap-org/9303003. Mora, T., Bialek, W., 2011. Are biological systems poised at criticality? Journal of Statistical Physics 144, 268-302. Moretti, P., Muñoz, M.A., 2013. Griffiths phases and the stretching of criticality in brain networks. Nature communications 4. Mosqueiro, T.S., Maia, L.P., 2013. Optimal channel efficiency in a sensory network. Physical Review E 88, 012712. Murray, J.D., Bernacchia, A., Freedman, D.J., Romo, R., Wallis, J.D., Cai, X., Padoa-Schioppa, C., Pasternak, T., Seo, H., Lee, D., Wang, X.J., 2014. A hierarchy of intrinsic timescales across primate cortex. Nature neuroscience 17, 1661-1663. Nagel, K., Herrmann, H.J., 1993. Deterministic models for traffic jams. Physica A: Statistical Mechanics and its Applications 199, 254-269. Nakamura, T., Kiyono, K., Yoshiuchi, K., Nakahara, R., Struzik, Z.R., Yamamoto, Y., 2007. Universal scaling law in human behavioral organization. Physical review letters 99, 138103. Nakamura, T., Takumi, T., Takano, A., Aoyagi, N., Yoshiuchi, K., Struzik, Z.R., Yamamoto, Y., 2008. Of mice and men--universality and breakdown of behavioral organization. PLoS ONE 3, e2050. Nejad, A.B., Ebdrup, B.H., Siebner, H.R., Rasmussen, H., Aggernaes, B., Glenthoj, B.Y., Baare, W.F., 2011. Impaired in antipsychotic-naive patients with first-episode schizophrenia. The world journal of biological psychiatry : the official journal of the World Federation of Societies of Biological Psychiatry 12, 271-281. temporoparietal deactivation with working memory load Nelson, T.S., Suhr, C.L., Freestone, D.R., Lai, A., Halliday, A.J., Mclean, K.J., Burkitt, A.N., Cook, M.J., 2011. Closed-loop seizure control with very high frequency electrical stimulation at seizure onset in the GAERS model of absence epilepsy. International journal of neural systems 21, 163-173. Nevado-Holgado, A.J., Marten, F., Richardson, M.P., Terry, J.R., 2012. Characterising the dynamics of EEG waveforms as the path through parameter space of a neural mass model: application to epilepsy seizure evolution. Neuroimage 59, 2374-2392. Niedermeyer, E., Sherman, D.L., Geocadin, R.J., Hansen, H.C., Hanley, D.F., 1999. The burst- suppression electroencephalogram. Clinical EEG (electroencephalography) 30, 99-105. Nielson, D.M., Smith, T.A., Sreekumar, V., Dennis, S., Sederberg, P.B., 2015. Human hippocampus represents space and time during retrieval of real-world memories. Proceedings of the National Academy of Sciences 112, 11078-11083. Nykter, M., Price, N.D., Larjo, A., Aho, T., Kauffman, S.A., Yli-Harja, O., Shmulevich, I., 2008. Critical networks exhibit maximal information diversity in structure-dynamics relationships. Physical review letters 100, 058702. Palva, J.M., Palva, S., 2011. 22 Roles of multiscale brain activity fluctuations in shaping the variability and dynamics of psychophysical performance. Progress in brain research 193, 335. Palva, J.M., Zhigalov, A., Hirvonen, J., Korhonen, O., Linkenkaer-Hansen, K., Palva, S., 2013. Neuronal long-range temporal correlations and avalanche dynamics are correlated with behavioral scaling laws. Proceedings of the National Academy of Sciences 110, 3585-3590. Papanikolaou, S., Bohn, F., Sommer, R.L., Durin, G., Zapperi, S., Sethna, J.P., 2011. Universality beyond power laws and the average avalanche shape. Nature Physics 7, 316-320. Penny, W.D., 2012. Comparing dynamic causal models using AIC, BIC and free energy. Neuroimage 59, 319-330. Perković, O., Dahmen, K., Sethna, J.P., 1995. Avalanches, Barkhausen noise, and plain old criticality. Physical review letters 75, 4528. 37 Criticality in the brain Cocchi et al. 38 Petermann, T., Thiagarajan, T.C., Lebedev, M.A., Nicolelis, M.A., Chialvo, D.R., Plenz, D., 2009. Spontaneous cortical activity in awake monkeys composed of neuronal avalanches. Proceedings of the National Academy of Sciences of the United States of America 106, 15921- 15926. Petzschner, F.H., Weber, L.A., Gard, T., Stephan, K.E., 2017. Computational Psychosomatics and Computational Psychiatry: Towards a joint framework for differential diagnosis. Biological Psychiatry. Plenz, D., Thiagarajan, T.C., 2007. The organizing principles of neuronal avalanches: cell assemblies in the cortex? Trends in neurosciences 30, 101-110. Ponce-Alvarez, A., He, B.J., Hagmann, P., Deco, G., 2015. Task-Driven Activity Reduces the Cortical Activity Space of the Brain: Experiment and Whole-Brain Modeling. PLoS computational biology 11, e1004445. Priesemann, V., Valderrama, M., Wibral, M., Le Van Quyen, M., 2013. Neuronal avalanches differ from wakefulness to deep sleep--evidence from intracranial depth recordings in humans. PLoS computational biology 9, e1002985. Priesemann, V., Wibral, M., Valderrama, M., Propper, R., Le Van Quyen, M., Geisel, T., Triesch, J., Nikolic, D., Munk, M.H., 2014. Spike avalanches in vivo suggest a driven, slightly subcritical brain state. Frontiers in systems neuroscience 8, 108. Publio, R., Ceballos, C.C., Roque, A.C., 2012. Dynamic range of vertebrate retina ganglion cells: Importance of active dendrites and coupling by electrical synapses. PloS one 7, e48517. Rabinovich, M., Volkovskii, A., Lecanda, P., Huerta, R., Abarbanel, H., Laurent, G., 2001. Dynamical encoding by networks of competing neuron groups: winnerless competition. Physical review letters 87, 068102. Rabinovich, M.I., Huerta, R., Varona, P., Afraimovich, V.S., 2008. Transient cognitive dynamics, metastability, and decision making. PLoS computational biology 4, e1000072. Razi, A., Kahan, J., Rees, G., Friston, K.J., 2015. Construct validation of a DCM for resting state fMRI. Neuroimage 106, 1-14. Ribeiro, T.L., Copelli, M., Caixeta, F., Belchior, H., Chialvo, D.R., Nicolelis, M.A., Ribeiro, S., 2010. Spike avalanches exhibit universal dynamics across the sleep-wake cycle. PLoS ONE 5, e14129. Rice, J., Ruina, A., 1983. Stability of steady frictional slipping. Journal of applied mechanics 50, 343- 349. Roberts, D.C., Turcotte, D.L., 1998. Fractality and self-organized criticality of wars. Fractals 6, 351- 357. Roberts, J.A., Boonstra, T.W., Breakspear, M., 2015. The heavy tail of the human brain. Current opinion in neurobiology 31, 164-172. Roberts, J.A., Friston, K.J., Breakspear, M., 2017. Clinical applications of stochastic dynamic models of the brain, part II: A review. Biological Psychiatry: Cognitive Neuroscience and Neuroimaging. Roberts, J.A., Iyer, K.K., Finnigan, S., Vanhatalo, S., Breakspear, M., 2014a. Scale-free bursting in human cortex following hypoxia at birth. The Journal of Neuroscience 34, 6557-6572. Roberts, J.A., Iyer, K.K., Vanhatalo, S., Breakspear, M., 2014b. Critical role for resource constraints in neural models. Frontiers in systems neuroscience 8, 154. Roberts, J.A., Perry, A., Lord, A.R., Roberts, G., Mitchell, P.B., Smith, R.E., Calamante, F., Breakspear, M., 2016. The contribution of geometry to the human connectome. Neuroimage 124, 379-393. Roberts, J.A., Wallis, G., Breakspear, M., 2013. Fixational eye movements during viewing of dynamic natural scenes. Frontiers in psychology 4, 797. Robinson, P., Rennie, C., Rowe, D., 2002. Dynamics of large-scale brain activity in normal arousal states and epileptic seizures. Physical Review E 65, 041924. Romani, G.L., Williamson, S.J., Kaufman, L., 1982. Tonotopic organization of the human auditory cortex. Science 216, 1339-1340. Rubinov, M., Sporns, O., Thivierge, J.P., Breakspear, M., 2011. Neurobiologically realistic determinants of self-organized criticality in networks of spiking neurons. PLoS computational biology 7, e1002038. 38 Criticality in the brain Cocchi et al. 39 Rubinov, M., Sporns, O., van Leeuwen, C., Breakspear, M., 2009. Symbiotic relationship between brain structure and dynamics. BMC neuroscience 10, 55. Ruderman, D.L., Bialek, W., 1994. Statistics of natural images: Scaling in the woods. Physical review letters 73, 814-817. Scarpetta, S., de Candia, A., 2013. Neural avalanches at the critical point between replay and non- replay of spatiotemporal patterns. PLoS One 8, e64162. Scheffer, M., Bascompte, J., Brock, W.A., Brovkin, V., Carpenter, S.R., Dakos, V., Held, H., Van Nes, E.H., Rietkerk, M., Sugihara, G., 2009. Early-warning signals for critical transitions. Nature 461, 53-59. Schrauwen, B., Büsing, L., Legenstein, R. (2009) On computational power and the order-chaos phase transition in reservoir computing. In: 22nd Annual conference on Neural Information Processing Systems (NIPS 2008). pp. 1425-1432. NIPS Foundation. Schuster, H.G., Plenz, D., Niebur, E., 2014. Criticality in neural systems. John Wiley & Sons. Scott, G., Fagerholm, E.D., Mutoh, H., Leech, R., Sharp, D.J., Shew, W.L., Knöpfel, T., 2014. Voltage imaging of waking mouse cortex reveals emergence of critical neuronal dynamics. The Journal of Neuroscience 34, 16611-16620. Sethna, J.P., Dahmen, K.A., Myers, C.R., 2001. Crackling noise. Nature 410, 242-250. Shew, W., 2015. Neuronal Avalanches. Encyclopedia of Computational Neuroscience, 2018-2024. Shew, W.L., Clawson, W.P., Pobst, J., Karimipanah, Y., Wright, N.C., Wessel, R., 2015. Adaptation to sensory input tunes visual cortex to criticality. Nature Physics 11, 659-663. Shew, W.L., Plenz, D., 2013. The functional benefits of criticality in the cortex. The neuroscientist 19, 88-100. Shew, W.L., Yang, H., Yu, S., Roy, R., Plenz, D., 2011. Information capacity and transmission are maximized in balanced cortical networks with neuronal avalanches. Journal of neuroscience 31, 55-63. Shew, W.L., Yang, H.D., Petermann, T., Roy, R., Plenz, D., 2009. Neuronal Avalanches Imply Maximum Dynamic Range in Cortical Networks at Criticality. Journal of Neuroscience 29, 15595-15600. Shih, H.-Y., Hsieh, T.-L., Goldenfeld, N., 2015. Ecological collapse and the emergence of travelling waves at the onset of shear turbulence. Nature Physics. Shriki, O., Alstott, J., Carver, F., Holroyd, T., Henson, R.N., Smith, M.L., Coppola, R., Bullmore, E., Plenz, D., 2013. Neuronal avalanches in the resting MEG of the human brain. Journal of Neuroscience 33, 7079-7090. Sreekumar, V., Dennis, S., Doxas, I., 2016. The Episodic Nature of Experience: A Dynamical Systems Analysis. Cognitive Science. Sreekumar, V., Dennis, S., Doxas, I., Zhuang, Y., Belkin, M., 2014. The geometry and dynamics of lifelogs: discovering the organizational principles of human experience. PloS one 9, e97166. Stam, C.J., de Bruin, E.A., 2004. Scale-free dynamics of global functional connectivity in the human brain. Human brain mapping 22, 97-109. Stanley, H.E., 1987. Introduction to phase transitions and critical phenomena. Introduction to Phase Transitions and Critical Phenomena, by H Eugene Stanley, pp. 336. Foreword by H Eugene Stanley. Oxford University Press, Jul 1987. ISBN-10: 0195053168. ISBN-13: 9780195053166 1. Stanley, H.E., 1999. Scaling, universality, and renormalization: Three pillars of modern critical phenomena. Reviews of modern physics 71, S358. Stepp, N., Plenz, D., Srinivasa, N., 2015. Synaptic plasticity enables adaptive self-tuning critical networks. PLoS computational biology 11, e1004043. Stevens, S.S., 1975. Psychophysics. Transaction Publishers. Stewart, C.V., Plenz, D., 2006. Inverted-U profile of dopamine–NMDA-mediated spontaneous avalanche recurrence in superficial layers of rat prefrontal cortex. The Journal of neuroscience 26, 8148-8159. Stramaglia, S., Pellicoro, M., Angelini, L., Amico, E., Aerts, H., Cortés, J., Laureys, S., Marinazzo, the Human Connectome. arXiv preprint Ising Model on D., 2015. Conserved arXiv:1509.02697. 39 Criticality in the brain Cocchi et al. 40 Suffczynski, P., Kalitzin, S., Da Silva, F.L., 2004. Dynamics of non-convulsive epileptic phenomena modeled by a bistable neuronal network. Neuroscience 126, 467-484. Tagliazucchi, E., Balenzuela, P., Fraiman, D., Chialvo, D.R., 2012. Criticality in large-scale brain FMRI dynamics unveiled by a novel point process analysis. Frontiers in physiology 3, 15. Tagliazucchi, E., Chialvo, D.R., Siniatchkin, M., Amico, E., Brichant, J.-F., Bonhomme, V., Noirhomme, Q., Laufs, H., Laureys, S., 2016. Large-scale signatures of unconsciousness are consistent with a departure from critical dynamics. Journal of The Royal Society Interface 13, 20151027. Thornton, T.L., Gilden, D.L., 2005. Provenance of correlations in psychological data. Psychonomic Bulletin & Review 12, 409-441. Tognoli, E., Kelso, J.A., 2014. The metastable brain. Neuron 81, 35-48. Tononi, G., Sporns, O., Edelman, G.M., 1994. A measure for brain complexity: relating functional segregation and integration in the nervous system. Proceedings of the National Academy of Sciences 91, 5033-5037. Touboul, J., Destexhe, A., 2010. Can power-law scaling and neuronal avalanches arise from stochastic dynamics. PLoS ONE 5, e8982. Touboul, J., Destexhe, A., 2015. Power-law statistics and universal scaling in the absence of criticality. arXiv preprint arXiv:1503.08033. Van de Ville, D., Britz, J., Michel, C.M., 2010. EEG microstate sequences in healthy humans at rest reveal scale-free dynamics. Proceedings of the National Academy of Sciences of the United States of America 107, 18179-18184. Van Kessenich, L.M., de Arcangelis, L., Herrmann, H., 2016. Synaptic plasticity and neuronal refractory time cause scaling behaviour of neuronal avalanches. Scientific reports 6, 32071. Van Orden, G.C., Holden, J.G., Turvey, M.T., 2005. Human cognition and 1/f scaling. Journal of Experimental Psychology: General 134, 117. Vano, J., Wildenberg, J., Anderson, M., Noel, J., Sprott, J., 2006. Chaos in low-dimensional Lotka– Volterra models of competition. Nonlinearity 19, 2391. Vasa, F., Shanahan, M., Hellyer, P.J., Scott, G., Cabral, J., Leech, R., 2015. Effects of lesions on synchrony and metastability in cortical networks. Neuroimage. Virkar, Y.S., Shew, W.L., Restrepo, J.G., Ott, E., 2016. Feedback control stabilization of critical dynamics via resource transport on multilayer networks: How glia enable learning dynamics in the brain. Physical Review E 94, 042310. Vojta, T., Avella, A., Mancini, F. (2013) Phases and phase transitions in disordered quantum systems. In: AIP Conference Proceedings. pp. 188-247. AIP. Vuong, Q.H., 1989. Likelihood ratio tests for model selection and non-nested hypotheses. Econometrica: Journal of the Econometric Society, 307-333. Wagenmakers, E.-J., Farrell, S., Ratcliff, R., 2004. Estimation and interpretation of 1/fα noise in human cognition. Psychonomic Bulletin & Review 11, 579-615. Wendling, F., Bartolomei, F., Bellanger, J., Chauvel, P., 2002. Epileptic fast activity can be explained by a model of impaired GABAergic dendritic inhibition. European Journal of Neuroscience 15, 1499-1508. Wendling, F., Hernandez, A., Bellanger, J.-J., Chauvel, P., Bartolomei, F., 2005. Interictal to ictal transition in human temporal lobe epilepsy: insights from a computational model of intracerebral EEG. Journal of Clinical Neurophysiology 22, 343. Whitfield-Gabrieli, S., Thermenos, H.W., Milanovic, S., Tsuang, M.T., Faraone, S.V., McCarley, R.W., Shenton, M.E., Green, A.I., Nieto-Castanon, A., LaViolette, P., Wojcik, J., Gabrieli, J.D., Seidman, L.J., 2009. Hyperactivity and hyperconnectivity of the default network in schizophrenia and in first-degree relatives of persons with schizophrenia. Proceedings of the National Academy of Sciences of the United States of America 106, 1279-1284. Yang, C.-N., Lee, T.-D., 1952. Statistical theory of equations of state and phase transitions. I. Theory of condensation. Physical Review 87, 404. Yang, D.-P., Zhou, H.-J., Zhou, C., 2017. Co-emergence of multi-scale cortical activities of irregular firing, oscillations and avalanches achieves cost-efficient information capacity. PLoS computational biology 13, e1005384. 40 Criticality in the brain Cocchi et al. 41 Yang, G.J., Murray, J.D., Repovs, G., Cole, M.W., Savic, A., Glasser, M.F., Pittenger, C., Krystal, J.H., Wang, X.-J., Pearlson, G.D., 2014. Altered global brain signal in schizophrenia. Proceedings of the National Academy of Sciences 111, 7438-7443. Yang, H.D., Shew, W.L., Roy, R., Plenz, D., 2012. Maximal Variability of Phase Synchrony in Cortical Networks with Neuronal Avalanches. Journal of Neuroscience 32, 1061-1072. Yeo, B.T., Krienen, F.M., Sepulcre, J., Sabuncu, M.R., Lashkari, D., Hollinshead, M., Roffman, J.L., Smoller, J.W., Zollei, L., Polimeni, J.R., Fischl, B., Liu, H., Buckner, R.L., 2011. The organization of the human cerebral cortex estimated by intrinsic functional connectivity. J Neurophysiol 106, 1125-1165. Zalesky, A., Fornito, A., Cocchi, L., Gollo, L.L., Breakspear, M., 2014. Time-resolved resting-state brain networks. Proceedings of the National Academy of Sciences of the United States of America 111, 10341-10346. Zapperi, S., Castellano, C., Colaiori, F., Durin, G., 2005. Signature of effective mass in crackling- noise asymmetry. Nature Physics 1, 46-49. Zapperi, S., Lauritsen, K.B., Stanley, H.E., 1995. Self-organized branching processes: mean-field theory for avalanches. Physical review letters 75, 4071. Zhigalov, A., Arnulfo, G., Nobili, L., Palva, S., Palva, J.M., 2017. Modular co-organization of functional connectivity and scale-free dynamics in the human brain. Network Neuroscience. 41 Criticality in the brain Cocchi et al. 42 Box 1: Ten-Point Summary 1. Criticality arises when a system is close to dynamic instability and is reflected by scale-free temporal and spatial fluctuations 2. Critical temporal fluctuations (crackling noise) occur in simple systems close to a bifurcation 3. Critical spatiotemporal fluctuations (avalanches) occur in complex systems close to a phase transition 4. Crackling noise and avalanches have now been observed in a wide variety of neuronal recordings, at different scales, in different species, and in health and disease 5. Computational models suggest a host of adaptive benefits of criticality, including maximum dynamic range, optimal information capacity, storage and transmission and selective enhancement of weak inputs 6. Resting-state EEG and fMRI data show evidence of critical dynamics 7. The onset of a specific cognitive function may reflect the stabilization of a particular subcritical state under the influence of sustained attention 8. Mounting evidence and models suggest that several neurological disorders such as epilepsies and neonatal encephalopathy reflect bifurcations and phase transitions to pathological states 9. Novel insights into neuropsychiatric disorders such as schizophrenia and melancholia might also be obtained by leveraging the tools of criticality, although this currently remains somewhat speculative 10. While the application of criticality to neuroscience is an exciting field, progress needs to proceed with due caution, using appropriate methods, considering alternative complex processes and using computational models in partnership with data analysis 42 Criticality in the brain Box 2: Glossary Cocchi et al. 43 The attractor of a dynamical system is the set of all points traversed once initial transients have passed. Attractors can be fixed points (with steady state solutions), limit cycles (periodic) or chaotic (deterministic but dynamically unstable and aperiodic). The basin of attraction of an attractor is the set of all initial conditions that eventually flow onto that attractor. A system can have more than one attractor even if all its parameters are fixed. Such a system is said to be multistable. Such systems will also have multiple basins of attraction, separated by basin boundaries. Bistability occurs in a multistable system which has exactly two attractors. A metastable system does not have any attractors. It instead has a series of saddles (fixed points with attracting and repelling subspaces) that are linked into a complex (heteroclinic) cycle. A metastable system will jump endlessly from the neighborhood of one saddle to another. Metastable systems are also called winnerless competition. An attractor that only changes slightly (and smoothly) when its underlying parameters are changes is said to be structurally stable. If the topology (shape) of the attractor fundamentally changes then the attractor is said to be unstable – or, alternatively undergo a bifurcation. The value of the parameter at which that discontinuous change occurs is said to be a critical or bifurcation point. A system consisting of many interacting components may exhibit a sudden change in state in the presence of a slowly tuned control parameter (such as temperature). Such a transition is called a phase transition and typically separates an ordered from a disordered state. Technically, a phase transition corresponds to a discontinuity in the thermodynamic free energy of a system. Criticality occurs when a system is poised at the point of a dynamic instability. Because of this, microscopic fluctuations are not damped but instead appear at all scales of the system. This yields power-law fluctuations in the temporal domain ("crackling noise") and the spatiotemporal domain ("avalanches"). A critical system will show scaling laws, such that a single (universal) function can map the shape of fluctuations at any scale into those at the scale above (or below). Bifurcations may be super-critical (when stable oscillations appear above the critical point) or sub- critical (when a zone of bistability occurs below the critical point). Only supercritical bifurcations can yield power law (critical) fluctuations: Sub-critical bifurcations lead to multistable switches that occur on a characteristic time scale. Phase transitions may be continuous (second order) or discontinuous (first order). Mathematically, these are equivalent to a super- and sub-critical bifurcations, respectively. As with bifurcations, critical power law fluctuations only occur in the neighbourhood of a continuous phase transition. Discontinuous phase transitions can yield multistable switching or complex mixtures of states (such as in boiling water). 43 Criticality in the brain Cocchi et al. 44 −𝑘+1 ) for the same critical exponent k. Some systems need to be externally tuned by a control parameter close to their critical point. In other systems, criticality will emerge from many initial parameter values, usually due to plasticity and memory. The order parameter of a complex system is a macroscopic observable such as the magnetic field of a ferromagnet. A non-zero order parameter arises in the ordered state of a system, in the supercritical (or active) phase. In the subcritical phase, the order parameter remains at zero even with the addition of energy. Such a state is called an absorbing state. A power law exists between two variables x and y if they obey the relationship: 𝑦 ∝ 𝑥−𝑘. If the power law arises in the setting of criticality, then the constant k is called the critical exponent. A system shows power law behaviour if the probability density function of its fluctuations obeys 𝑓(𝑥) ∝ 𝑥−𝑘 for all values of x greater than some minimum cut-off 𝑥𝑚𝑖𝑛. A power law probability distribution is also called a Pareto distribution. The corresponding cumulative distribution obeys 𝑃𝑟𝑜𝑏(𝑋 > 𝑥) ∝ 1 − (𝑥 ⁄ 𝑥𝑚𝑖𝑛 A system is said to be scale-free when it doesn't have a characteristic time or length scale. Scale-free systems show a power law probability distribution over several orders of magnitude with an exponent k that is less than 2. Correspondingly, the variance of a scale-free system is only bounded by the system size. A complex system shows evidence of slowing down when it is close to a critical point – that is, the time scale of its fluctuations (the characteristic return to the mean) slow down, changing from a fast (exponential) process to a slow power law. A system has an exponential distribution when its probability density function is given by 𝑓(𝑥) ∝ 𝑒−𝑥/𝐿. The system has a single time scale, corresponding to the constant exponent L. Any distribution whose probability distribution drops off more slowly than an exponential distribution is said to be heavy-tailed. The Pareto distribution 𝑓(𝑥) ∝ 𝑥−𝑘 is a classic heavy tailed distribution but log-normal and stretched exponential (Weibull) distributions are also heavy-tailed. Many physical systems close to a critical point do have some weak damping that acts on the largest fluctuations: These systems show an exponentially-truncated power law, 𝑓(𝑥) ∝ 𝑥−𝑘𝑒−𝑥𝑙. 44
1903.10626
1
1903
2019-03-25T23:06:22
The role of physiological complexity changes in resting-state EEG in clinical effectiveness of rTMS and tDCS in treatments of resistant depression
[ "q-bio.NC" ]
The present literature about possible mechanisms behind the effectivity of noninvasive electromagnetic stimulation in major depressive disorder (MDD) is not very rich. Despite extensive research in applications for clinical practice, the exact effects are yet not clear. We are comparing our previous results about the complexity changes induced by repetitive Transcranial Magnetic Stimulation (rTMS), and transcranial direct current stimulation (tDCS) which are known to modulate neural dynamics. Also, we are reviewing different biomarkers of complexity changes connected to depression, and how they change with the stimulation. TDCS is low-intensity TES, known to have polarity specific effects (neuromodulatory effects), and rTMS is inducing an electric field in the tissue circumstantially via Faraday's law. Both nonlinear modalities of electromagnetic stimulation may affect the levels of physiological complexity in the brain. We also compare the changes of complexity in electroencephalogram (EEG) and electrocardiogram (ECG), as potential future predictors of therapy outcome.
q-bio.NC
q-bio
The role of physiological complexity changes in resting-state EEG in clinical effectiveness of rTMS and tDCS in treatments of resistant depression 1Department for General Physiology and Biophysics, University of Belgrade, Belgrade, Serbia 2Amsterdam Health and Technology Institute, HealthInc, Amsterdam, the Netherlands Milena Čukić1,2, PhD Abstract The present literature about possible mechanisms behind the effectivity of noninvasive electromagnetic stimulation in major depressive disorder (MDD) is not very rich. Despite extensive research in applications for clinical practice, the exact effects are yet not clear. We are comparing our previous results about the complexity changes induced by TMS, and tDCS which are known to modulate neural dynamics. Also, we are reviewing different biomarkers of complexity changes connected to depression, and how they change with the stimulation. TDCS is low-intensity TES, known to have polarity specific effects (neuromodulatory effects), and rTMS is inducing an electric field in the tissue circumstantially via Faraday's law. Both nonlinear modalities of electromagnetic stimulation may affect the levels of physiological complexity in the brain. We also compare the changes of complexity in electroencephalogram (EEG) and electrocardiogram (ECG), as potential future predictors of therapy outcome. Keywords: Transcranial direct current stimulation (tDCS), repetitive Transcranial Magnetic Stimulation (rTMS), Electroencephalogram (EEG), Variability Heart Rate (VHR), treatment-resistant depression, Anodal stimulation, Cathodal stimulation, plasticity, Theta burst stimulation (TBS), Neuromodulation, Neurostimulation, Rehabilitation, Major Depressive Disorder (MDD) Introduction The course of depression is very often recurrent and can become chronic with highly likely relapse rates within one year of remission from 35% to 80% (Eaton et al., 2008; Fekadu et al., 2009). Usual treatments for depression are pharmacological and psychological, but Rush and colleagues report that one third fail to reach remission (Rush et al., 2006). Non-invasive electromagnetic modalities of stimulation, like transcranial magnetic stimulation (TMS) and transcranial electric stimulation (TES) are alternatives to this usual practice. We are going to briefly review several application modalities of TMS and low-intensity modality of TES, namely tDCS. Here we want to focus on elucidating the effectiveness of two electromagnetic treatment considered 'novel.' My putting it under the quotation marks is intentional because transcranial magnetic stimulation was launched as a commercial stimulator 1985 (Prof. Antony Baker, 1985) and it cannot be still novel after almost 35 years of research and diagnostic applications. Similar reasoning applies to tDCS which is low-intensity stimulation modality of transcranial electrical stimulation (TES), but still, some details important for an understanding of its mechanisms are missing. Like Opitz put it in recent research, the important point is in interpretability of TES effects (Opitz et al., 2018): 'if electric fields are delivered inconsistently, but effects are observed nevertheless, the results are more difficult to interpret because effect could be driven by other incidentally affected brain regions.' TMS was initially introduced as a non-invasive tool for investigating and mapping cortical functioning and connectivity (Barker et al., 1985). TMS primarily use a strong magnetic field to induce an electric field in the surface layers of the brain (cortex) and without pain characteristic to ECT, initiate optimally focused activation of neural structures. One of its modalities used in psychiatry is repetitive TMS (rTMS). While standard TMS induce single (or sometimes) paired pulses, repetitive TMS delivers repeated pulses showed to enable more extended modulation of neural activity. It can be slow (if repetitions are up to 1Hz/one per second) or fast (with 10Hz or higher frequencies). Fast rTMS have excitatory, while slow rTMS applications have inhibitory effects (Rosa and Lisanby, 2012). It is interesting to note that rTMS was finally confirmed for reimbursements in a treatment-resisting depression just last year (2018) in the Netherlands, 34 years from its début in medicine (US Food and Drug Administration- FDA approved the first rTMS device for treatment of MDD in 2008). The basis for rTMS treatment is a poor response to at least one pharmacological agent in the current episode (O'Reardon et al., 2007). We are going to elucidate on that question 'how long it takes for a scientifically proven innovation to be translated to clinical practice?' The scientific literature about a safe and useful application of TMS is so rich that it is hard to believe that it took for so long. The reason for using TMS in treatments of depression came from a demonstration of functional impairments in prefrontal cortices and limbic regions (Atkinson et al., 2014) and very well based frontal asymmetry in alpha (Allen et al., 2004). There are several different modalities of rTMS use. Since high-frequency stimulation can be uncomfortable during the initial stimulation period, low-frequency rTMS was introduced to minimize headaches and scalp discomfort; it is also lowering the risk for developing seizures (Rossi et al., 2009). When the stimulation is ongoing over the left and right dorsolateral prefrontal cortex-DLPFC (simultaneously, or stimulation one side after another), it is called Bilateral application of rTMS. Conca and colleagues argue that this kind of application may reinstate any imbalance in prefrontal neural activity (Conca et al., 2002). Fitzgerald showed that likelihood for a clinical response increases by providing both types of stimulation (Fitzgerald et al., 2006). Deep TMS (dTMS) can stimulate larger brain volumes and deeper structures (Roth et al., 2007) which is maybe more relevant to the pathophysiology of depression (Costafreda et al., 20013; Kwaasteniet et al., 2013; Atkinson et al., 2014). Theta burst stimulation (TBS) (Huang et al., 2005) utilize high and low frequencies in the same stimulation train; it delivers three bursts at high frequency (50Hz) with an inter-burst interval of 5Hz (theta range) frequency. Two different approaches to the application of TBS exists: continuous and intermittent (cTBS and iTBS). cTBS delivers 300 to 600 pulses without interruption, and iTBS delivers 30 pulses (trains) every 10s during 190s (600 pulses in total) (Chuang et al., 2015). While cTBS reduces cortical excitability, iTBS increases it; they are mimicking the process of long term potentiation (LTP) and long term depression (LTD) (Stagg and Nitsche, 2016; Huang et al., 2005). LTP allows for modulation of synaptic strength that stabilizes for days, months, or even years and has therefore been postulated as a likely candidate for memory formation in the brain (Anderson and Lomo, 1966; Bliss and Lomo, 1973). LTD has been studied extensively in the hippocampus and refer to highly specific processes; synaptic plasticity with very similar properties has been demonstrated in the neocortex and is therefore commonly referred to as LTP-like plasticity (Stagg and Nitsche, 2016). The main advantage of TMS is the reduced time of treatment in comparison to all other rTMS modalities. For conventional rTMS it takes 20-45 min, while for TMS it is typically less than 5 min; also it operates at 80% of the resting motor threshold (MT) and may be more comfortable than rTMS. Another modality of stimulation synchronized TMS (sTMS), is a new treatment paradigm which delivers stimulation synchronized to an individual's alpha frequency (Jin and Phillips, 2014). It is based on rTMS ability to make entrainment of oscillatory activity to the programmed frequency of stimulation with the aim of restoring normal oscillatory activity (Leuchter et al., 2013). It does not cause neural depolarization; therefore it may cause fewer adverse effects. In opposition to all modalities mentioned above of TMS, tDCS which is low-intensity modality of TES has specific advantage of use; its low cost and portability are opening the possibilities of use out of the clinical setup. tDCS is not triggering action potentials directly (unlike TMS), and it is shown that it modulates cortical excitability by shifting neural membrane resting potential (Nitsche et al., 2008). TDCS effects can outlast the stimulation period up to two hours (Nitsche, 2001) and even show significant effects for more extended periods. In our recent research, we showed that tDCS is able of causing the shift in the brain-state (Čukić et al., 2018) based on EEG recordings reconstruction in phase-space, for at least half an hour. The primary advantage of tDCS compared to any TMS is its ease of administration, portability, much less recorded adverse events (AE) and lower price. Mutz and colleagues aimed at examining the efficacy and acceptability of non-invasive brain stimulation in adult unipolar and bipolar depression (Mutz et al., 2018). They found the most substantial evidence for high-frequency rTMS over left DLPFC, followed by low- frequency rTMS over the right DLPFC and bilateral rTMS. Intermittent TBS provides a potential advance in terms of reduced treatment duration. Authors also stated that tDCS is a potential treatment for non-resistant depression which has demonstrated efficacy in terms of response as well as remission (Mutz et al., 2018). Tuomas Neuvonen and colleagues started making their anytime anywhere tDCS applications in Soma already. They elaborated on interindividual differences as one of the major sources of misinterpretation (Fonteneau et al., 2018) Meta-analyses on the efficacy of rTMS and tDCS in treatment-resistant depression In the recent tDCS review study, authors discussed the interaction of induced electric field (EF) with tissue, and on electroporation and concluded that the intensities needed for electroporation remain orders of magnitude above tDCS (Antal et al., 2017). Typical current in tDCS montage is 1-2 mA (0.03-2mA/cm2 current to electrode area ratios depending on the electrode size) which results in cortical EF strengths up to 0.4-0.8 V/m (Ruffini et al., 2013b). Both the applied current and the resting brain EFs are ~1000 times lower than those for pulsed stimulation used in electroconvulsive therapy (Alam et al., 2016) and are considered to be below the required intensity for evoking an action potential in resting cell (Radman et al., 2009). They are still capable of modifying the spontaneous firing rates of neurons, and also of inducing plasticity processes in neural networks (Ranieri et al., 2012; Jackson et al., 2016). Neural networks generate their own characteristic EFs in intracellular spaces, and this kind of stimulus can influence their molecular and structural changes (Frohlich and McCormick, 2010). Based on the work of Mutanen and Ilmoniemi on the induced brain-shift in an application of TMS, we later aimed at elucidating the effect of tDCS by utilization of the same method. Frohlich and McCormick detected the alteration of transmembrane potential for 0.5-1.3 mV, with the intensity of the induced electric field of 2-4 mV/m. Both Miranda and Saturnino predicted values of the maximum EF in the cortex of realistic head model are between 0.2 V/m and 0.5 V/m (if the current is 1mA) (Miranda et al., 2013; Saturnino et al., 2015). According to Nelson and Nunneley, tDCS induced energy in the cortex is 0.1 mW/kg, for the current of 1mA, a conductivity of 0.4 S/m and the median value of EF 0.5 V/m (Nelson and Nunneley, 1998). Antal and other experts concluded in their 2017 guidelines that those values caused by tDCS are safe (Antal et al., 2017). Wagner reviewed similar details important to understand the electromagnetic interaction involved from previous literature (Wagner et al., 2007). In case of TMS, a magnetic field (~1-4T pulsed over 0.125-1ms dependent of parameters) induces an electric field that drives currents in the brain of a magnitude approximately 5.13x10-8 A/m in the cortex per 1A/s steady-state source. Currents are carried by free charges and ions (known as ohmic or resistive currents, or volume conduction) or through the depolarization of dipoles in the tissue layers (Wagner et al., 2007). Fregni and colleagues were searching for predictors of antidepressant response in clinical trials of transcranial magnetic stimulation (Fregni et al., 2006). At the time they did their research, it was already known that rTMS has a significant antidepressant effect, the results were heterogeneous. They hypothesized that individual patients' characteristics might contribute to such heterogeneity in a sample comprising of 195 patients. Results showed that age and treatment refractoriness were significant negative predictors of depression improvement when adjusted to other significant predictors. The findings of Mutz and colleagues (2018) are in line because they also found that younger people are more probable to be responders. Another study on the effectivity 'Transcranial direct current stimulation in treatment-resistant depression: a randomized double-blind, placebo-controlled study' was published in 2012 (Palm et al., 2012). They concluded that 'anodal tDCS, applied for two weeks, was not superior to placebo treatment in patients with treatment-resistant depression. However, secondary outcome measures are pointing to a positive effect of tDCS on emotions'. The authors recommended modified and improved tDCS protocols to be carried out in controlled pilot trials to develop tDCS towards an effective antidepressant intervention in therapy-resistant depression. Berlim and colleagues published two studies on the same topic in 2013 and 2014 (Berlim et al., 2013, 2014). The first one was about the efficiency of tDCS, and the other on the efficiency of rTMS. They first concluded that the clinical utility of tDCS as a treatment for MDD remains unclear when clinically relevant outcomes such as response and remission rates are considered. Moreover, they recommended future studies on larger samples. In their 2014 study, they concluded that 'Meta-analyses have shown that high-frequency (HF) repetitive transcranial magnetic stimulation (rTMS) has antidepressant properties when compared with sham rTMS.' However, its overall response and remission rates in major depression (MD) remain unclear (Berlim et al., 2014). Bennabi and his colleagues concluded in their 2015 study that 'tDCS did not induce the clinically relevant antidepressant effect in active and sham stimulation groups. There was no impact on psychomotor and neuropsychological functioning' (Bennabi et al., 2015). So, as Fregni pointed out, the results are still heterogeneous. Brunoni and colleagues aimed to assess tDCS efficacy and to explore individual response predictors in their 2016 study (Brunoni et al., 2016). Since tDCS was extensively investigated for the treatments, particularly of major depression, they performed another review study by utilization of individual patient data (contrary to aggregate data approach in many review studies before). The current literature confirms that tDCS can induce neuromodulatory changes in cortical activity, and essential for depression, to ameliorate depressive symptoms (Brunoni et al., 2012). Its antidepressant effects are based on previous findings of hypoactivity of the left dorsolateral prefrontal cortex (DLPFC). Brunoni concentrated on providing precise estimates of tDCS efficacy based on depression improvement, response and remission rates. They tried to identify variables associated with tDCS efficacy (Brunoni et al., 2016). They also used several up-to-date software tools for detecting the bias, to re-analyze the original data and so on. After careful literature research and selection (among 153 studies), their review study comprises of six studies, only (they included studies which are double-blinded, sham-controlled, reporting on a number of repetitions, dosage, medication, and other detailed data). However, on end among those six studies were three (half of all selected) studies published by the same group which is also a bias (Loo et al., 2010; Loo et al., 2012; Brunoni et al., 2013). The primary dependent variable was clinical response since the authors claim that it has better performance over remission which presented heterogeneity of cut-offs. They used a number needed to treat (NNT) as a measurement which illustrates clinical intervention effectiveness; the higher the NNT, the less effective the intervention. The authors found significant effect sizes of the efficacy of active vs. sham tDCS in terms of depression improvement response and remission, with overall response and remission rates of active tDCS of 34% and 23% respectively (NNT of 7 and 9) (Brunoni et al., 2016). The authors further compared their own meta-analysis study (on six studies, half of which are their own) with a Cochrane meta- analysis assessing the efficacy of antidepressant drug treatments in primary care (Arrol et al., 2009) finding their results are in line. Brunoni and colleagues also claim their findings to be in the same range as another meta- analysis for depression which utilized active vs. sham repetitive transcranial magnetic stimulation (rTMS) that found a response and remission NNTs of 6 and eight respectively. So, for treatment response and remission rTMS meta-analysis study(Berlim et al, 2014) found absolute rates of 29.3% and 18.6%, while tDCS meta-analysis found 34% and 23% respectively. The only problem here was that out of 289 patients included in tDCS study there were just 147 patients with an active stimulation, and in rTMS study, the sample is almost ten-fold; n=1371. From the point of statistical learning, the other sample would be expected to give us reliable results, while the first sample for this type of comparison is considered to be very modest. Brunoni and colleagues also recommended two potential applications of tDCS in treatments for depression: in primary care settings and as a non-pharmacological, neuromodulatory therapy for depression (Brunoni et al., 2016). In another meta-analytical study performed by Antal et al. (2017) the authors commented on limitations of Brunoni study, and used a much larger sample, covering the publications and data collected from approximately 8000 people. They concluded that tDCS is considered safe, giving guidelines for avoiding Adverse Events (AE) as well as questionnaires for further research (designed to prevent all previously reported AEs). In their meta-analysis comprising of 56 previously published studies, Mutz and colleagues found about actual effectivity of different modalities of TMS used for treatment-resistant depression (Mutz et al., 2018). They reported that high-frequency rTMS (over left DLPFC) was associated with improved rates of response as well as remission in comparison with sham treatment. Only one of the reviewed studies recruited bipolar depression patients (Nahas et al., 2003) but with demonstrated antidepressant efficacy. They also found that low-frequency rTMS over the right DLPFC was associated with significantly higher response and remission rates than sham stimulation. Low-frequency rTMS over lDLPFC did not show significant improvement (there were no significant differences in response rates compared to sham). Bilateral rTMS demonstrated significant improvement in response but not remission rates compared to sham. When dTMS is concerned, it also showed significant improvement (for both response and remission rates), but response rates were marginally higher relative to sham. Neither response nor remission rates for sTMS were significantly higher than a sham. Intermittent TBS over the left DLPFC was associated with a fivefold improvement in response rates compared to sham. No evidence was found of improvement in case of cTBS over rDLPFC and bilateral TBS. In case of tDCS, they confirmed significant improvement in both response and remission rates in comparison with sham; their analysis suggested tDCS to be effective only in patients with non-treatment resistant depression (Mutz et al., 2018). In contrast to most rTMS trials, tDCS showed a potential initial therapeutic option for depression. The evidence that rTMS and tDCS are activating deeper structures A study performed by Li and colleagues provides a valuable insight into mechanisms by which tDCS may modulate cognitive function and also has implications for the design of future stimulation studies (Li et al., 2018). They showed that the effects of tDCS on brain activity are dependent on the cognitive state. Both anodal and cathodal tDCS produced widespread BOLD changes in brain areas anatomically remote from the critical area being stimulated. Those results are in line with Sven Bestmann's findings, which first demonstrated that even with the sub- threshold stimulation of TMS, deep structures are activated (Bestmann et al., 2003 and 2004). It may come as a surprise, but we are looking here at the evidence that both tDCS and rTMS are affecting other structures than intended. It may be that it is in connection with the finding that severe depression does have the problem with functional connectivity, due to the deep white matter tracts significant in a fronto-limbic system (Kwaasteniet et al., 2013; Kim et al., 2013). The focus of a majority of researchers was for so long at defining the polarity specific and other noticed effects, but the problem is that both methods are noninvasive and therefore only direct measurement can be a solution for this riddle. Opitz et al (Opitz et al., 2015, 2016, 2018) and Alekseichuk (Alekseichuk et al., 2018) applied the robust physical approach in elucidating this complex interaction between the induced electrical field and brain tissue. Opitz performed the direct measurement from implanted electrodes (in direct empirical validation of TES) in patients who were operated in a procedure intended to implant a grid of stimulation electrodes for their epilepsy, and compared the results with the cohort of 25 healthy people (neurotypical individuals) from connectome project. The motivation for their work was the lack of exact measurements and the problem with variability between individuals. Opitz and colleagues combined direct intracranial measurements of electric fields generated by TES in surgical epilepsy patients with computational modeling (Opitz et al., 2018). They directly validated the computational models and identified key parameters needed for accurate model predictions. They also derived practical guidelines for a reliable application of TES in terms of the precision of electrode placement needed to achieve a desired electric field distribution. When comparing the measured and predicted values of induced electrical fields, they encounter the problem of conductivity constants values. They are reported in the literature with great variety, so further research is needed (Opitz et al., 2018). To establish standards in this field would be of significant importance to increase the reliability of future stimulation protocols. When they were testing the model developed for the study they found that maximum strength of the induced field predicted by the model was larger than that they directly measured. The mean electric field strength was 0.058 mV/mm for the measurement results, while the mean electric field strength over the best fitting stimulation results was 0.100 mV/mm (range 0.071-0.122 mV/mm) (Opitz et al., 2018). In another patient with a different implanted grid of electrodes, the trend was in the opposite direction; mean measured field strength was 0.115 mV/mm, while predicted was 0.060 mV/mm. The authors concluded that computational models using current standard conductivities slightly misestimate the measured field estimate strength, suggesting the need for individual adjustments of conductivities (Huang et al., 2017). They also demonstrated the importance of factors like the accuracy of the skull or surgical component modeling; other factors as gyral folding and CSF thickness showed to have a profound effect on TES induced fields (Opitz et al., 2015). They urge other researchers to direct their efforts to directly measure conductivities to improve realistic modeling, or even include more tissue types (Aydin et al., 2014). It is based on their finding that the variation of the position of an electrode may maximally vary to 1cm, due to folding in the brain surface (electricity concentrate differently on sharp and flat surfaces, depending on geometry) in order to provide efficient stimulation. In their work, it is also confirmed how important it is to accurately model the shape of electrodes for realistic field calculations (Saturnino et al., 2015). In another study (Alekseichuk et al., 2018) they systematically evaluated the induced electrical fields and analyzed their relationship to brain anatomy (TMS and TES induced fields were compared in finite element method, in a mouse, capuchin monkey, and human model). The theoretical explanation and final calculation of the induced electrical field during the stimulation was necessary for those who are further developing computational models. Results from our previous research about the changes of complexity induced by the application of TMS and tDCS In my Doctoral dissertation, and before that in my magisterial thesis, I explored the parameters influencing the strength of stimulation (Čukić, 2006) and the changes in complexity due to TMS (Čukić, 2011). I am going to report here some of our results which can help understand the changes that both TMS and tDCS can induce, based on co-recording of the electrophysiological signal. Now, if we want to understand how a stimulus is acting upon a system, we can observe the levels of complexity before the stimulation, and after the stimulation and by comparing those levels we can conclude about the influence. This is precisely what we did with two mentioned modalities of stimulation. In our work from 2011/12, we compared single pulse TMS induced complexity changes with the complexity changes induced by peripheral electrical stimulation (the same target muscle was FDI), and we showed that single pulse TMS is able to decrease the complexity of the system (Čukić et al., 2013). In the first set of results, we aimed at examining the changes in surface electromyogram (EMG) recorded from the first dorsal interosseous muscle (FDI) which was the target muscle in TMS protocol. In that protocol, we were also changing the level of voluntary contraction to test how the amount of present force in the muscle can interfere with the effect of stimulation. On the first graph from those results (Figure 1) we can see surface EMG recorded in that experiment and also magnified epochs of that signal we used for further fractal analysis. Figure 1: The raw electromyogram signal recorded from the target muscle (FDI) in a TMS stimulation protocol. The epochs for analysis extracted from the trace 'before' and 'after' the stimulation are magnified. They were used for further analysis to probe for the complexity changes induced by the stimulation (Čukić et al., 2013). After comparative classical spectral and fractal analysis, we found the significant changes in the epochs recorded in participants (all healthy persons) before and after the stimulation. The following two illustrations are showing both the spectral and fractal result of the analysis. Figure 2: The distributions of frequnecies (spectra) before (left) and after (right) stimulation in the surface EMG signal; the change induced by the stimulus are not immediately obvious (Čukić et al., 2010). Figure 3: Amplitude spectra of surface EMG before (left) and after (right) the TMS stimulation. The difference in complexity levels can be seen with a naked eye (Ćukić et al., 2010). It can be seen from both graphs that the drop in complexity of the signal is apparent. The level of complexity in a healthy system (being a brain, a heart or a muscle) is showing systems' ability to adapt. When an influence out of the dynamical system impact the system, the change can be observed via changes in complexity (or irregularity). In this case, it can be seen that after the application of TMS, the complexity of surface myogram reduced. One can see that on the following figures: 4 and 5. The first is showing a histogram, constructed from fractal dimensions calculated from sections of EMG signal, showing the distribution of its values. The red line compared to black dotted one is showing how those FDs calculated from epochs of EMG before and after the stimulation are distributed. Those representing 'after' sections are shorter, meaning that the facilitation happened during this stimulation. Another figure (4) is representing the mean values of calculated FD (from EMG) showing a significant difference between them. Figure 4: The changes in Higuchi fractal dimension calculated from surface EMG epochs before (red) and after the stimulation. The distribution is slightly shifted to the left, illustrating the shortened latencies in the response after the stimulation (Čukić, 2006). Figure 5: The average values of Higuchi Fractal dimension before (dotted line) and after the stimulation; it can be seen that the complexity of signal dropped after the stimulation (Čukić, 2012). Several years after that experiment, our colleagues from Finland published a study which examined the change TMS induced in brain state (in phase-space) of 10 healthy participants (Mutanen et al., 2013). They applied recurrence plot analysis on EEG concurrently recorded with TMS, and they also stimulated motor cortex (the same target muscle as we did in our experiment). This is why I believe that our results are in a way comparable and compatible. They showed that even some time after the stimulation is over, the system (brain) stayed in a highly improbable and higher energy state in comparison to the resting state before the stimulation. The following diagram is illustrating the process they described in their work (Mutanen et al., 2013), on our graphical representation similar to theirs (Figure 6). Figure 6: A) The blue arrow represents the primary current source, a flow of ions in synapses in the motor cortex. The dashed blue line represents returning volume current. Jp is the primary current density (induced by the stimulation). B) This schematic representation shows how TMS affects brain-state. The green and red curves are pre- and post- TMS trajectories (of a brain state). The spontaneous activity (green) occupy one region in state-space. After the stimulation, it occupies another region (red) which is in a higher energy state. Gradually, over time, the system will return to lower energy and more probable state. The projections of those two states are in EEG signal space, i and j standing for two different channels. In the signal, space trajectories are measured only at discrete time points (dotted projections, red and green). The green line is representing the trajectory of a system before the stimulation, and red line the trajectory in another part of phase space after the TMS. Both trajectories have the projections on the plane (dots in the plane Ei Ej), which are the samples of recorded EEG (Ilmoniemi and Kičić, 2010). Jp and Jv are the current densities illustrating the ongoing process during the stimulation; Jp is causing the movement of charges in the tissue (it appears due to the induced electrical field in the tissue), and Jv is volume conductor density current which is a counterpart process to the first one. In essence, they showed that due to the stimulation the system is staying in elevated state (the brain-shift is confirmed by the measures of Global recurrence, MSS and SV) long enough so we can detect it. Next thing we did, was to test the same methodology, but on another type of stimulation, tDCS. We are presenting here part of the results which are dealing with complexity changes. On figure 6, it can be seen how HFD and SampEn detected the changes in complexity after tDCS stimulation. We know that this kind of stimulation is very different, due to the much lower level of energy delivered to the tissue, and we know that it is considered to have just modulatory character. In this research (in 2015) we tried to test the changes in complexity before and after the stimulation ('before' was considered a resting state since we did not have sham to compare to, and after was immediately after-t1, or a half an hour after-t2). However, if you can detect changes that survive more than half an hour later, does that imply early plastic-like changes? Unfortunately, we were not able to confidently interpret the results, because contrary to our previous knowledge about just modulatory effect of tDCS we detected temporary decrease and after that increase of complexity measured by Higuchi fractal dimension and Sample entropy. What is more, it looked as at some combinations there is no difference between cathodal and anodal stimulation, in terms of complexity changes. Due to the small sample, we opted to apply another methodology set, but are working on repeating with the same method in presently ongoing research. Nitsche showed much earlier that induced changes in system survive more than two hours (Nitsche, 2001). It would be interesting to see whether the complexity stays on a different level after a couple of days, or a week or even longer. Figure 7: Comparison of complexity changes before and after anodal (left) and cathodal (right) stimulation, represented with two calculated measures Higuchi fractal dimension and sample entropy. The epochs extracted from EEG traces (on ten electrodes) were before the simulation (t0), immediately after the stimulus (t1) and half an hour after the stimulation (t2) (Čukić, 2017). After all those examples, we can say that both TMS and tDCS are capable of changing the complexity of the system (brain). Although the mechanism and power of their stimulation are different even more than order (power) of the value of the induced field, some similarities can be drawn from our results. If we know that in depression an elevated complexity on cortex is observed, and that majority of researchers agree that this may serve as a biomarker for depression, is it too big a stretch if we think that maybe both rTMS and tDCS can help because they are both diminishing that complexity? If a significant number of people suffering from depression is reporting amelioration of symptoms (let us remember that remission is defined as 'symptoms becoming bearable') they might profit from the ability of electromagnetic stimulation to at least temporarily, decrease the problematic elevated excitability in their DLPFC and elsewhere. We have to stress here, that as well as in ECT, both tDCS and rTMS are requiring maintenance therapy since the benefit reported by patients is lasting for a limited timeframe. To conclude, further research is needed, on a much larger sample. Are rTMS and tDCS capable of changing the complexity of the brain? In our previous work, we were interested in complexity changes in the human brain after electromagnetic stimulation (Čukić et al., 2001, 2008, 2013, 2018). The logic behind this is that the complex system (and the brain is one of the most complex dynamical systems that we know) inherently exhibit certain levels of complexity with a possible purpose of accommodation to internal or external changes of parameters which affect the conditions and surroundings of the system is operating in. We have to be aware that we can observe this only indirectly. We are analyzing an EEG as a product of the complex dynamical system (EEG is a composite signal comprising of individual electric signals as a consequence of activation of different neurons and neuronal groups). An electrical signal generated by neural networks when superimposed give us the record from every point we positioned an electrode on the scalp. Of course, we can say that EEG is an electrical representation of brain functioning, but what it is showing us are the voltages picked up from the brain surface (cortex). We can only speculate how that signal represents the deeper structures activities. Despite the extensive use in research and treatment, our knowledge about exact mechanisms by which tDCS is influencing different structures in the brain is limited. Lucia Li and her colleagues examined how transcranial direct current stimulation modulates brain network function (Li et al., 2018). They used MRI for simultaneous recording whit tDCS. From resting state recordings, the main effect of tDCS was to accentuate default mode network (DMN) activation and salience network (SN) deactivation. In contrast, during task performance, tDCS increased SN activation (Li et al., 2018). In the absence of a task, the main effect of anodal tDCS was more pronounced, whereas cathodal tDCS had a more significant effect during task performance. Cathodal tDCS also accentuated the within DMN connectivity associated with the performance. There were minimal main effects of stimulation on network connectivity. These results demonstrate that right inferior frontal gyrus tDCS can modulate the activity and functional connectivity of large-scale brain networks involved in cognitive function, in a brain state and polarity dependent manner. Bestmann showed that with sub-threshold stimulation, MRI captured the activation of the auditory system, transversal and superior temporal girys, inferior colliculus and mediate geniculate nucleus, which can be understood as change induced by synaptic transmission (induction of changes of excitability in sensory-motor areas, M1/S1) (Bestmann et al., 2003). When we apply any artificial influence (like artificially induced electric field in proximity of sensitive neural or muscle tissue) which we know is efficient in inducing action potential (or smaller changes in electrical sense) in the tissue it could be observed through the lens of nonlinear analysis of composite signal (i.e. EEG or EMG)(Čukić, 2006, 2011 and 2013). The levels of complexity induced by this artificial (un-natural) stimulus are changed due to the electrical nature of the phenomena (Čukić, 2011). If we are using TMS we are inducing an artificial electrical field (via Faraday's law) which in turn force electrical charges to move through the tissue (Wagner et al., 2008; Ridding and Rothwell, 2007; Thickbroom, 2007; Bestmann, 2008). When we try to compare those two stimulations (rTMS and tDCS), the data from the literature are very diverse. We must say that the fundamental biophysical, or physics research followed the advent of TMS just shortly (Paton and Amassian, 1987). Harris and Miniussi elucidated the effects of stimulation in cognitive experimenting focusing on the mechanism of action of induced electrical field on cognitive performances (Harris and Miniussi, 2008), and above mentioned authors explored those details in another, more physiological or biophysical level (like Bestmann, 2008). Both electromagnetic techniques are evolving, and we have to invent better, safer and deeper penetrating noninvasive solutions for depressive disorders. We would like once again to return to the connection between the electromagnetic stimulation in cases of depression and dynamic of heart rhythm (ECG); we already elaborated on a connection between variability heart rate and response on ECT, or potential forecasting of responders based on nonlinear measures. Researchers who explored the connection of VHR and outcome of therapy concluded that '…low baseline HRV is associated with rapid relapse of depression after ECT. Both high baseline HRV and increasing HRV predict a sustained outcome' (Karpyak et al., 2004). Other researchers in the field also noticed that the low level of VHR corresponds with the severity of disease and that VHR often increases when a patient reaches the remission (Bozkurt et al, 2013). So, without repeating a summary of many physiological complexity studies in this particular task, the conclusion is that the dynamics of the heart is characteristically aberrated in persons diagnosed with bipolar or unipolar depression, and also in burnout syndrome. Lower levels of complexity in heart rhythm are also indicating a potential risk of developing cardiovascular diseases in people with depression. In comparison with EEG analysis results in depression, we can say that the detection from two physiological signals (EEG and ECG) has the opposite trend, for physiological reasons. In EEG studies, an elevated complexity in depressive patients can be observed and measured (and used as potential neuromarker). In ECG studies, a lower level of nonlinear measures (as well as classical ones) and generally lower complexity can be observed in depressives. However, both of them can be used as useful biomarkers to early detect and monitor the developments in every single case. The difference is that recording of EEG requires a visit to the clinic, while the recording of ECG is already possible with portable monitoring devices. A similar parallel exists between the utilization of rTMS and tDCS; the latter can be used in primary care and even at home, which is a much more accessible option to patients and their families. Like DST was used to differentiate melancholic depression from other less severe depression forms (Shorter and Fink, 2010), so fractal and nonlinear measures of EEG (or ECG) can serve as useful biomarkers necessary for the improvement of accuracy of clinical diagnostic and treatment. Let us return for a moment to one of the results from our 2016 pilot study; we showed that there is a difference (which can be measured) between the EEG complexity of patients diagnosed with depression who are in acute episode and those who already are in remission. Counterintuitively, those who remitted, exhibited higher levels of complexity, in comparison with participants who were still in exacerbation (Čukić et al., 2018). This phenomenon still needs further exploration. We are currently working on collecting more data, because we believe that even an opposite trend could be present if we screen patients later, after the remission (after possible neural reorganization). Some results from studies dealing with the effect of mindfulness- based cognitive therapy (MBCT) showed that during eight weeks of treatment (or training), the connection between the insula and ventromedial prefrontal cortex were uncoupled, as a marker of improvement measurable within the fronto-limbic system (Davidson et al., 2004; Slagther et al., 2007). It would be interesting to explore further how applications of different measures of complexity on both electrophysiological signals (EEG or ECG) can improve mental healthcare practice. They could become a valuable tool in decision making for clinicians. When the interaction between electromagnetic stimulation and medication is concerned, we know much more about that, than about the changes of complexity or irregularity, due to a much richer body of evidence in existing literature. For example, among other guidelines, it is recommended in potential tDCS use (Antal et al., 2017) that specific clinical presentation (for example bipolar depressive disorder) requires some mood stabilizers, and not some of the drugs which might attenuate the anodal effect of tDCS. I will add here one anecdotal example as an illustration of that interaction. In the beginning of our project on Mental Healthcare institution, during our conversation about the possibilities of detecting epileptic foci by utilization of a graph theory method based on high density EEG, a colleague mentioned, that it maybe correspond with the fact that in the last five years many patients who are suffering from depression are actually having excellent reaction on anticonvulsants. Is it possible that they have a focus in their limbic system? We discarded that idea as highly unrealistic, but after the finalizing results of several projects on that topic, I think that they are reacting well because of their elevated excitability, induced by their decreased functional connectivity in frontolimbic system. The same can apply to the effectiveness of rTMS and tDCS in cases of treatment-resisting depression; maybe they are effective because of their already demonstrated ability to decrease the levels of complexity in their systems? We also know that the solution is alas, of a temporary nature. As patients reacting well on electroconvulsive therapy need to have later maintenance ECT, it is also required in other forms of electromagnetic stimulation. Both rTMS and tDCS last for a certain amount of time. They are not capable of resolving the problem, and they can help improve the patient's state for a while. The effect is also individual; whatever mode of stimulation we consider, we see that intricate technical details (like the position of electrodes, the physical characteristics of a stimulator), or individual differences (like the thickness of a cortex, or a bone, or different conductivity of a tissue) play their role in final effect of the treatment. Another advantage of the therapeutic use of electromagnetic stimulation is, that opposite to any medication they do not have side effects; or at least they certainly cannot be (or are so far away of being) life-threatening. Hence, this kind of detection (by utilization of nonlinear measures of electrophysiological signals) are, according to our opinion, one of the prerequisites of a personal medicine in mood disorders. Conclusions Asgdsfg Acknowledgements We want to express our gratitude to our colleagues from Italy Professor Carlo Miniussi, Debora Brignani, Ph.D. and Maria Concetta Pelliciari, Ph.D. for sharing their EEG recordings (data are from the Pellicciari et al., 2013. paper) with us and for all valuable discussions. References Eaton, W.W. et al. (2008). Population-based study of first onset and chronicity in major depressive disorder. Arch. Gen. Psychiatry, 65, 513-520. Fekadu A., et al., (2009) What happens to patients with treatment-resistant depression? A systematic review of medium to long term outcome studies. Journal of Affective Disorders, 116, 4-11.Rush et al., 2006 Barker, A.T., Jalinous, R., Freeston, I.L. (1985). Non-invasive magnetic stimulation of human motor cortex. Lancet, 325, 1106-1107. Opitz, A., Paulus, W., Will, S., Antunes, A., Thielscher, A. (2015). Determinants of the electric field during transcranial direct current stimulation. Neuroimage, 109, 140 -- 150. doi: 10.1016/j.neuroimage.2015.01.033 Opitz, A., Fox, M.D., Craddock, R.C., Colcombe, S., Milham, M.P., 2016b. An integrated framework for targeting functional networks via transcranial magnetic stimulation. Neuroimage 127, 86 -- 96. https://doi.org/10.1016/j.neuroimage.2015.11.040.29. Opitz, A., Yeagle, E., Thielscher, A., Schroeder, C., Mehta, A. D, Milham, M.P. (2018). On the importance of precise electrode placement for targeted transcranial electric stimulation. Neuroimage, 181, 560-567. doi: 10.1016/j.neuroimage.2018.07.027. Rosa, M. A., Lisanby, S. H. (2012). Somatic treatment for mood disorders. Neuropsychopharmacology, 37, 102-116. O'Reardon, J.P., et al. (2007). Efficacy and safety of transcranial magnetic stimulation in the acute treatment of major depression: a multisite randomized controlled trial. Biological Psychiatry, 62, 1208-1216. Attkinson, L., Sankar, A., Adamas, T.M., Fu, C.H. (2014). Recent advances in neuroimaging of mood disorders:structural and functional neural correlates of depression, changes with therapy, and potential for clinical biomarkers. Current Treatment Options in Psychiatry, 1, 278-293.Allen, J. J. B., Urry, H. L., Hitt, S. K., Coan, J. A. (2004). The stability of resting frontal electroencephalographic asymmetry in depression. Psychophysiology, 41(2), 269 -- 280. http://dx.doi.org/10.1111/j.1469-8986.2003.00149.x. Rossi, S., Hallet, M., Rossini, P.M., Pascual-Leone, A., Group S.o.T.C. (2009). Safety, ethical consideration and application guidelines for the use of transcranial magnetic stimulation in clinical practice and research. Clin. Neurophysiology, 120, 2008-2039. Conca, A., et al. (2002). Combining high and low frequencies in rTMS antidepressive treatment: preliminary results. Hum. Psychopharmacol.: Clin. Exp. 17, 353-356. Fertonani, A., and Miniussi, C. (2017). Transcranial electrical stimulation: what we know and do not know about mechanisms. Neuroscientist, 23, 109 -- 123. doi: 10.1177/1073858416631966 Fitzgerald P.B., Borwn T.L., Marston NAU, Daskalakis ZJ, de Castella A, Kulkarni J. (2003) Transcranial magnetic stimulation in the treatment of depression: a double-blind, placebo-controlled trial. Arch Gen Psychiatry 60: 1002-1008. Fitzgerald, P.B., et al., (2006). A randomized, controlled trial of sequential bilateral repetitive transcranial magnetic stimulation for treatment-resistant depression. American Journal of Psychiatry, 163, 88-94. Roth, Y., Amir, A., Levkovitz, Y., Zangen, A. (2007), Three-dimensional distribution of the electric field induced in the brain by transcranial magnetic stimulation using figure-8 and deep H-coils. J. Clin. Neurophysiol, 24, 31-38. Costafreda, S.G. et al. (2013). Modulation of amygdala response and connectivity in depression by serotonin transporter polymorphism and diagnosis. Journal of Affective Disorders, 150, 96-103. de Kwaasteniet, B., Ruhe, E., Caan, M., et al. (2013). Relation Between Structural and Functional Connectivity in Major Depressive Disorder. Biological Psychiatry, 74(1), 40 -- 47. http://dx.doi.org/10.1016/j.biopsych.2012.12.024. Huang, Y.Z., Edwards, M.J., Rounis, E., Bhatia, K.P., Rothwel, J.C. (2005). Theta burst stimulation of the human motor cortex. Neuron, 45, 201-206. Chung, S., Hoy, K., Fitzgerald, P. (2015). Theta-burst stimulation: a new for of TMS treatment for depression? Depress. Anxiety, 32, 182-192. Stagg, C. J., Nitsche, M. A. (2011). Physiological Basis of Transcranial Direct Current Stimulation. The Neuroscientist, 17(1), 37-53. Anderson P, Lomo T. 1966. Mode of activation of hippocampal pyramidal cells by excitatory synapses on dendrites. Exp Brain Res 2(3):247 -- 60. Bliss TV, Lomo T. 1973. Long-lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. J Physiol 232(2):331 -- 56. Jin, Y., Phillips, B. (2014). a pilot study of the use of EEG-based synchronized Transcranial Magnetic Stimulation (sTMS) for treatment of major depression. BMC Psychiatry, 14, 1. Lefaucheur, J. P., Andre-Obadia, N., Antal, A., Ayache, S. S., Baeken, C., Benninger, D. H., et al. (2014). Evidence-based guidelines on the therapeutic use of repetitive transcranial magnetic stimulation (rTMS). Clinical Neurophysiology, 125, 2150-206. Nitsche, M. A., Cohen, L. G., Wassermann, E. M., Priori, A., Lang, N., Antal, A., Paulus, W., Hummel, F., Boggio, P.S., Fregni, F., Pascual-Leone, A. (2008). Transcranial direct current stimulation: State of the art 2008. Brain Stimulation, 1(3), 206-23. doi: 10.1016/j.brs.2008.06.004. Epub 2008 Jul 1. Nitsche, M. A., Fricke, K., Henschke, U., Schlitterlau, A., Liebetanz, D., Land, N., Hening, S., Terau, F., Paulus, W. (2003). Pharmacological modulation of cortical excitability shifts induced by transcranial direct current stimulation in humans. Journal of Neurophysiology, 553, 293-301. Mutz J., Edgcumble D.R., Brunoni A.R., Fu C.H.Y. (2018) Efficacy and acceptability of non-invasive brain stimulation for the treatment of adult unipolar and bipolar depression: A systematic review and meta-analysis of randomized sham-controlled trials. Neuroscience and Biobehavioral Reviews 92 (2018) 291-303. Fonteneau C. et al., (2018) Frontal Transcranial Direct Current Stimulation Induces Dopamine Release in the Ventral Striatum in Human. Cerebral Cortex, Volume 28, Issue 7, July 2018, Pages 2636 -- 2646, https://doi.org/10.1093/cercor/bhy093 Antal A., Alekseichuk I., Bikson M. et al., (2017) Low intensity transcranial electric stimulation: Safety, ethical, legal regulatory and application guidelines. Clin Neurophysiol. 2017 Sep; 128(9): 1774 -- 1809.Ruffini et al., 2013b Alam M, Truong DQ, Khadka N, Bikson M. 2016. Spatial and polarity precision of concentric high- definition transcranial direct current stimulation (HD-tDCS). Physics in Medicine and Biology 61:4506 -- 4521. doi: 10.1088/0031-9155/61/12/4506, PMID: 27223853 Radman T., Datta A., and Bikson M. (2009) One-dimensional representation of a neuron in a uniform electric field. Conf Proc IEEE Eng Med Biol Soc. 2009 ; 2009: 6481 -- 6484. doi:10.1109/IEMBS.2009.5333586. Ranieri F, Podda MV, Riccardi E, Frisullo G, Dileone M, Profice P, Pilato F, Di Lazzaro V, Grassi C.(2012). Modulation of LTP at rat hippocampal CA3-CA1 synapses by direct current stimulation. Journal of Neurophysiology 107:1868 -- 1880. doi: 10.1152/jn.00319.2011, PMID: 22236710 Frohlich F, McCormick DA. (2010). Endogenous electric fields may guide neocortical network activity. Neuron 67:129 -- 143. doi: 10.1016/j.neuron.2010.06.005, PMID: 20624597 Miranda PC, Pajevic S, Pierpaoli C, Hallett M, Basser PJ. (2001). The distribution of currents induced in the brain by magnetic stimulation: a finite element analysis incorporating DT-MRI-derived conductivity data. Proceedings of the International Society for Magnetic Resonance in Medicine 9. Miranda, P. C., Lomarev, M., Hallet, M. (2006). Modeling the current distribution during transcranial direct current stimulation. Clinical Neurophysiology, doi:10.1016/j.clinph.2006.04.009 Nitsche MA, Paulus W. (2001). Sustained excitability elevations induced by transcranial DC motor cortex stimulation in humans. Neurology 57:1899 -- 1901. doi: 10.1212/WNL.57.10.1899 Saturnino, G. B., Antunes, A., Thielscher, A. (2015). On the importance of electrode parameters for shaping electric field patterns generated by tDCS. Neuroimage, 120,25-35. Pelliciari, M. C., Brignani, D., Miniussi, C. (2013). Excitability modulation of the motor system induced by transcranial direct current stimulation: a multimodal approach. Neuroimage, 83, 569-580. Nelson D.A. and Nunneley S.A. (1998) Brain temperature and limits on transcranial cooling in humans: quantitative modeling results. Eur J Appl Physiol (1998) 78: 353. https://doi.org/10.1007/s004210050431 Wagner T., Valero-Cabre A., and Pascual-Leone A. (2007) Noninvasive Human Brain Stimulation. Annu. Rev. Biomed. Eng. 2007. 9:527 -- 65. Fregni, F., Marcolin, M.A., Mychowski, M., Amaz, R., Hasey, G., Rumi, D.O., et al. (2006). Predictors of antidepressant response in clinical trials of transcranial magnetic stimulation. International Journal of Neuropharmacology, 9, 641-54. Palm, U., Schiler, C., Fintescu, Z., Obermeier, M., Keeser, D., Reisinger, E. et al. (2012). Transcranial direct current stimulation in treatment resistant depression: a randomized double-blind, placebo-controlled study. Brain stimulation, 5, 242-51. Berlim, M.T., Van den Eynde, F., Daskalakis, Z.J. (2013). Clinical utility of transcranial direct current stimulation (tDCS) for treating major depression: a systematic review and meta-analysis of randomized, double- blinded and sham-controlled trials. J Psychiatr Res, 47, 1-7. Berlim, M.T., van der Eynde, F., Tovar-Perdomo, S., Daskalakis, Z.J. (2014). Response, remission and drop- out rates following high-frequency repetitive transcranial magnetic stimulation 9rTMS) for treating major depression: a systematic review and meta-analysis of randomized, double-blind and sham-controlled trials. Psychol Med, 44, 225-39. Bennabi, D., Nicolier, M., Monnin, J., Tio, G., Vandel, P. et al. (2015). Pilot study of feasibility of the effect of treatment with tdcs in patietis suffering from treatment-resistant depression treated with escitalopram. Clinical Neurophysiology, 126, 1185-1189. Brunoni A.R. et al., (2016) Transcranial direct current stimulationfor acute major depressive episodes: meta- analysis of individual patient data. The British Journal of Psychiatry 2016, 208,1-10. Doi: 10.1192/bjp.bp.115- 164715. Brunoni, A. R., Nitsche, M. A., Bolognini, N., Bikson, M., Wagner, T., Merabet, L. et al. (2012). Clinical research with transcranial direct current stimulation (tDCS): challenges and future direction. Brain Stimulation, 5, 175-195. Loo C.L. et al., (2012) Transcranial direct current stimulation for depression; 3-wkke randomized, sham- controlled trial. Br. J. Psychiatry 2012; 200:52-9. Loo C.K. et al., (2010). A double-blind sham-controlled trial of transcranial direct current stimulation for the treatment of depression. Int J Neuropsychpharmacol. 2010; 13: 61-9. Arroll, B., Elley, C.R., Fishman, T., Goodyear-Smith, F.A., Kenealy, T., Balashki, G., et al. (2009). Antidepressants versus placebo for depression in primary care. Cochrane Database Syst Rev, 3, CD007954.Nahas et al., 2003 Li, L. M., Violante, I. R., Leech, R., Ross, E., Hampshire, A., Opitz, A., Rothwell, J. C., Carmichael, D. W, Sharp, D. J. (2018). Brain state and polarity dependent modulation of brain networks by transcranial current stimulation. Human Brain Mapping, 1-12. DOI: 10.1002/hbm.24420. Bestmann, S., Baudewig, J., Siebner, H.R., Rothwell, J.C., Frahm, J. (2003). Is functional magnetic resonance imaging capable of mapping transcranial magnetic cortex stimulation? Supplements to Clinical Neurophysiology, 56, 55-62. Bestmann, S., Baudewig, J., Siebner, H.R., Rothwell, J.C., Frahm, J. (2003). Subthreshold high-frequency TMS of human primary motor cortex modulates interconnected frontal motor areas as detected by interleaved fMRI-TMS. Neuroimage, 20, 1685-1696. DOI: 10.1016/j.neuroimage.2003.07.028 Bestmann, S., Baudewig, J., Siebner, H. R., Rothwell, J. C., Frahm, J. (2004). Functional MRI of the immediate impact of transcranial magnetic stimulation on cortical and subcortical motor circuits. European Journal of Neuroscience, 19, 1950-1962. de Kwaasteniet, B., Ruhe, E., Caan, M., et al. (2013). Relation Between Structural and Functional Connectivity in Major Depressive Disorder. Biological Psychiatry, 74(1), 40 -- 47. http://dx.doi.org/10.1016/j.biopsych.2012.12.024. Kim D, Bolbecker AR, Howell J et al. Disturbed resting state EEG synchronization in bipolar disorder: a graph-theoretic analysis (2013). NeuroImage Clin. 2013; 2: 414 -- 423. DOI: 10.1016/j.nicl.2013.03.007. Alekseichuk I., Mantell K., Shirinpour S., Opitz A. (2018) Comparative Modeling of Transcranial Magnetic and Electric Stimulation in Mouse, Monkey, and Human. bioRxiv preprint first posted online Oct. 13, 2018; doi: http://dx.doi.org/10.1101/442426. Huang, Y., Liu, A. A., Lafon, B., Friedman, D., Dayan, M., Wang, X., Bikson, M., Doyle, W.K., Devinsky, O, Parra, L.C. (2017) Measurements and models of electroc fields in the in vivo human brain during transcranial electric stimulation. eLife, 6:e18834. DOI: 10.7554/eLife.18834 Aydin, U., Vorwerk, J., Kupper, P., Heers, M., Kugel, H., Galka, A., Hamid, L., Wellmer, J., et al. (2014). Combining MEG and EEG for the reconstruction of epileptic activity using a calibration realistic volume conductor model. PlosOne, 9. Saturnino, G. B., Antunes, A., Thielscher, A. (2015). On the importance of electrode parameters for shaping electric field patterns generated by tDCS. Neuroimage, 120,25-35. Čukić et al., (2018). EEG machine learning with Higuchi's fractal dimension and Sample Entropy as features for detection of depressive disorder. https://arxiv.org/abs/1803.05985 Čukić, M., Kolar, D., Kajkut, S. (2001). Primena Transkranijalne magnetne stimulacije u tretmanu depresije. [Application of Transcranial Magnetic Stimulation in treatments of Major Depression]. ME2.2, ETRAN 2001, pp. 231-234. Čukić, M., Kalauzi, A., Ljubisavljević, M. (2006). Uticaj rastojanja stimulacionog kalema i voljne aktivacije na latencije mišićnih odgovora na Transkranijalnu magnetnu stimulaciju [The Influence of distance of stimulation coil and voluntary activation on MEP latencies in TMS study]. ETRAN 2006, Beograd, book 3 ME2, str. 252-255. Čukić, M., Ljubisavljević, M., Kalauzi, A. (2009): Analysis of probability density distributions of TMS induced physiological parameters, Clinical neurophysiology, Vol 120.March 2009. pg e129.doi:10.1016/j.clinph.2008.12.013 Čukić, M., Kalauzi, A., Ilić, T., Mišković, M., Ljubisavljević, M. (2009). The influence of coil-skull distance on transcranial magnetic stimulation motor evoked responses, Experimental Brain Research, 192, 1, 53-60. Čukić, M., Filipović, S., Ljubisavljević, M. (2010). Transcranial magnetic stimulation effects on tremor in Primary writing tremor patients. First International Workshop on "Synaptic Plasticity: from bench to bed side" Taormina, Sicily, Italy, April 28.-May 1. 2010. Čukić, M., Platiša, M., Ljubisavljević, M., Kostić, V. (2012). Complexity Changes in TMS Induced Surface EMG. CCS2012, 4th International Interdisciplinary chaos symposium on chaos and complex systems, April 29- May 02. 2012, Antalya, Turkey. Čukić, M., Kalauzi, A., Ljubisavljevic, M., Jorgovanović, N., Kostić, V. (2012). The Evolution of Complexity in Transcranial Magnetic Stimulation induced surface EMG: a possible illustration of plasticity-like changes. The 16th Congress of the European Federation of Neurological Societies - EFNS 2012, Stockholm, September 8-11. 2012. Čukić, M., Oommen, J., Mutavdzic, D., Jorgovanovic, N., Ljubisavljevic, M. (2013). The effect of single- pulse transcranial magnetic stimulation and peripheral nerve stimulation on complexity of EMG signal: fractal analysis. Experimental Brain Research, 228(1), 97-104, DOI: 10.1007/s00221-013-3541-1. Čukić, M. (2006). The Influence of configuration and geometric characteristics of induced electrical field induced by transcranial magnetic stimulation on basic physiological parameters of excitability of motor cortex of man. Magisterium thesis, defended on March 16th 2006. Department for Physiology and Biophysics, University of Belgrade, School of Biology, Belgrade, Serbia. Mutanen, T., Nieminen, J. O., Ilmoniemi, R. (2013). TMS-evoked changes in brain-state dynamics quantified by using EEG data. Frontiers in Human Neuroscience, 25 April, 2013 Volume7, Article 155, doi: 10.3389/fnhum.2013.00155. Ilmoniemi, R., Kičić, D. (2010). Methodology for combined TMS and EEG. Brain Topogr. 22, 233- 248.Nitsche, 2001 Čukić, M., Pokrajac, D., Stokić, M., Simić, S., Radivojević, V., Ljubisavljević, M. (2018). EEG machine learning with Higuchi fractal dimension and Sample Entropy as features for successful detection of depression. Arxive.org/Preprint at Cornell repository for Statistics/Machine learning https://arxiv.org/abs/1803.05985. Thielscher A, Antunes A, Saturnino GB. (2015). Field modeling for transcranial magnetic stimulation: a useful tool to understand the physiological effects of TMS? Engineering in Medicine and Biology Society (EMBC), 2015, 37th Annual International Conference of the IEEE:IEEE. p. 222 -- 225. doi: 10.1109/EMBC.2015.7318340 Bestmann S.(2008) The physiological basis of transcranial magnetic stimulation. Trends in Cognitive Sciences. 12: 81-3. PMID 18243042 DOI: 10.1016/j.tics.2007.12.002 Amassian, V. E., Cracco, R. Q., Maccabee, P. J. (1989). Focal stimulation of human cerebral cortex with the magnetic coil: a comparison with electrical stimulation. Electroencephalogr. Clin Neurophysiol, 74, 401- 416. Amassian, V. E., Eberle, L., Maccabee, P. J., Cracco, R. Q. (1992). Modelling magnetic coil excitation of human cerebral cortex with a peripheral nerve immersed in a brain-shaped volume conductor: the significance of fiber bending in excitation, Electroencephalogr Clin Neurophysiol 85, 291 -- 301. Amassian, V. E., Stewart, M., Quirk, Q. J. (1987). Physiological basis of motor effects of a transient stimulus to the cerebral cortex. Neurosurgery, 20, 74-93. Wagner S, Rampersad SM, Aydin U , Vorwerk J, Oostendorp TF, Neuling T, Herrmann CS, Stegeman DF, Wolters CH. (2014). Investigation of tDCS volume conduction effects in a highly realistic head model. Journal of Neural Engineering 11:016002. doi: 10.1088/1741-2560/11/1/016002, PMID: 24310982 Harris J.A., Clifford C.W.G. and Miniussi C. (2008) The Functional Effect of Transcranial Magnetic Stimulation: Signal Suppression or Neural Noise Generation? MIT 2018, Journal of Cognitive Neuroscience 20:4, pp. 734 -- 740. Karpyak, V. M., Rasmussen, K. G., Hammill, S. C., Mrazek, D. A.(2004). Changes in heart rate variability in response to treatment with electroconvulsive therapy. .Journal of ECT 20(2):81-8. Bozkurt A, Barcin C, Isintas M, Ak M, Erdem M, Ozmenler KN. (2013) Changes in heart rate variability before and after ECT in the treatment of resistant major depressive disorder. Isr J Psychiatry Relat Sci. 2013;50(1):40-6. Shorter Edward and Fink Max (2010) Endocrine Psychiatry: Solving the Riddle of Melancholia. Oxford University Press, 2010. Davidson, R. J. (2004). What does the prefrontal cortex "do" in affect: perspectives on frontal EEG asymmetry research? Biological Psychology, 67(1 -- 2), 219 -- 234. http://dx.doi.org/10.1016/j.biopsycho.2004.03.008. Slagter HA, Lutz A, Greischar LL, Francis AD, Nieuwenhuis S, Davis JM, Davidson RJ.(2007) Mental training affects distribution of limited brain resources. Plos Biology. 5: e138. PMID 17488185 DOI: 10.1371/journal.pbio.0050138
1609.05730
1
1609
2016-09-19T13:56:01
Multi-contact synapses for stable networks: a spike-timing dependent model of dendritic spine plasticity and turnover
[ "q-bio.NC" ]
Excitatory synaptic connections in the adult neocortex consist of multiple synaptic contacts, almost exclusively formed on dendritic spines. Changes of dendritic spine shape and volume, a correlate of synaptic strength, can be tracked in vivo for weeks. Here, we present a combined model of spike-timing dependent dendritic spine plasticity and turnover that explains the steady state multi-contact configuration of synapses in adult neocortical networks. In this model, many presynaptic neurons compete to make strong synaptic connections onto postsynaptic neurons, while the synaptic contacts comprising each connection cooperate via postsynaptic firing. We demonstrate that the model is consistent with experimentally observed long-term dendritic spine dynamics under steady-state and lesion induced conditions, and show that cooperation of multiple synaptic contacts is crucial for stable, long-term synaptic memories. In simulations of a simplified network of barrel cortex, our plasticity rule reproduces whisker-trimming induced rewiring of thalamo-cortical and recurrent synaptic connectivity on realistic time scales.
q-bio.NC
q-bio
Multi-contact synapses for stable networks: a spike-timing dependent model of dendritic spine plasticity and turnover Moritz Deger1,2, Alexander Seeholzer1, Wulfram Gerstner1 1: School of Computer and Communication Sciences and School of Life Sciences, Brain Mind Institute, École Polytechnique Fédérale de Lausanne, 1015 Lausanne EPFL, Switzerland 2: Institute for Zoology, Faculty of Mathematics and Natural Sciences, University of Cologne, 50674 Cologne, Germany Abstract Excitatory synaptic connections in the adult neocortex consist of multiple synaptic contacts, almost exclusively formed on dendritic spines. Changes of dendritic spine shape and volume, a correlate of synaptic strength, can be tracked in vivo for weeks. Here, we present a combined model of spike-timing dependent dendritic spine plasticity and turnover that explains the steady state multi-contact cong- uration of synapses in adult neocortical networks. In this model, many presynaptic neurons compete to make strong synaptic connections onto postsynaptic neurons, while the synaptic contacts compris- ing each connection cooperate via postsynaptic ring. We demonstrate that the model is consistent with experimentally observed long-term dendritic spine dynamics under steady-state and lesion induced conditions, and show that cooperation of multiple synaptic contacts is crucial for stable, long-term synaptic memories. In simulations of a simplied network of barrel cortex, our plasticity rule reproduces whisker-trimming induced rewiring of thalamo-cortical and recurrent synaptic connectivity on realistic time scales. Excitatory synapses on pyramidal cells of the mammalian neocortex are almost exclusively made onto postsynaptic dendritic spines (see [1, 2]). Dendritic spines are small protrusions of the dendrite that vary in size and shape, and are subject to ongoing plasticity in both the developing and the adult brain [3, 4, 5, 6, 7, 8]. A central role of dendritic spines may be that the compartment of the spine allows for localized, synapse- specic plasticity mechanisms [1, 2]. Furthermore, plastic dendritic spines may allow neurons to select the subset of inputs from nearby axons of candidate presynaptic cells [9], so as to optimally process stimuli under the constraints of limited total tissue volume available for synaptic connections [10, 11]. Although dendritic spines show considerable volatility and may be formed and eliminated within a couple of days [12], some spines are maintained over long periods of time [12] and may support lifelong memories [13]. During learning, dendritic spines in the neocortex change selectively, and if potentiated spines are selectively shrunk by optogenetic means, learning is disrupted [14]. Time-lapse imaging of changes of spines on time-scales of seconds, hours and days to weeks [15, 12, 16, 4, 5, 17, 18] has shown that spine removal is tightly linked to the dynamics of the spine volume and to synaptic ecacy. Despite the ongoing dendritic spine turnover response patterns of neuronal networks can be long-term stable [19]. While past modeling studies of synaptic consolidation indicate how memories could be stored beyond the rst hour after plasticity induction [20, 21, 22, 23], spike-timing dependent plasticity (STDP) based models of synaptic plasticity [24, 25, 26] have not yet been linked quantitatively to dendritic spine plasticity and turnover [5], lifetime of synaptic contacts, and observed statistics of synaptic contact numbers [27, 28]. Synapses between pyramidal neurons in the somatosensory cortex of the rat are sparse and the distribution of the number of synaptic contacts for pairs of pre- and postsynaptic neurons has a characteristic bimodal shape [27, 28, 29]: most presynaptic neurons make no contact at all with a given postsynaptic neuron, but those that do are likely to establish 4 or more contacts [27]. However, the number of potential 1 synaptic contacts between a pair of neurons, based on the proximity of axons and dendrites in reconstructed neuronal morphologies, is always unimodally distributed [28, 30, 29] with a most likely value of 1 and an estimated mean value signicantly greater (Fig. 1D) than the mean number of actual contacts observed in experiments. This puzzling observation led to the hypothesis that the observed dierence between potential and actual contact numbers cannot arise from independent synapse formation, but instead requires a mechanism of cooperation between synaptic contacts onto the same neuron [28]. We wondered whether the cooperativity between synaptic contacts would necessarily imply a novel biochemical mechanism or whether known principles of STDP are sucient to explain the bimodal distribution of contact numbers in sparsely connected, stable neocortical networks. We propose a unifying approach to dendritic spine formation, plasticity and removal, in which STDP of single synaptic contacts provides the crucial cooperative mechanism regulating structural plasticity. Synaptic contacts connecting a neuron A to another neuron B share the same pre- and postsynaptic neuronal activity: presynaptic action potentials (spikes) cause Glutamate release and eventually leave a biochemical trace at each dendritic spine, while information about postsynaptic spikes may reach the same dendritic spines by means of backpropagating action potentials [31, 32].Since in some preparations spine plasticity [33, 34] and spine formation [18] are visible within less than a minute after stimulation, structural spine plasticity cannot be assumed to be slower than STDP, in contrast to previous models [35, 36, 37, 38, 39, 40]. Here, we present a quantitative model that links continuous synaptic plasticity of dendritic spines to discrete structural plasticity of synaptic contact formation and removal. We show that our model reproduces experimentally observed steady-state properties of dendritic spine plasticity and turnover in the somatosen- sory cortex of adult rodents. While existing theoretical models of the bimodal distribution of synaptic contacts in networks are algorithmic [28, 30], our model suggests that STDP is sucient to generate the bimodal distribution observed in experiments. Our model predicts that a synaptic contact that is part of a connection consisting of multiple synaptic contacts is more stable than an isolated synaptic contact. Moreover, a recurrent thalamo-cortical network model exhibits long-term stable structure despite ongoing plasticity, and lesion-induced rewiring on time scales of days and weeks. A combined model of STDP and structural plasticity In a rst set of simulated experiments, we model the plasticity of the synapses from N = 1000 presynaptic neurons onto a single postsynaptic neuron (Fig. 1A). All presynaptic model neurons re as independent Poisson processes with a rate of 5/s. A connection from a presynaptic neuron j to the postsynaptic cell may consist of several synaptic contacts k, with 1 ≤ k ≤ nj , where nj is the number of potential contacts [28, 30]. Each of these contacts is described by a unit-less weight wjk . The total weight of the connection from a presynaptic neuron j is given by the sum of the weights wjk over all its contacts k, wj =Pnj k=1 wjk . The contact weight wjk in our model describes how much the specic contact k contributes to ring the postsynaptic cell, and can be viewed as representing the dendritic spine volume which is in turn strongly correlated to the AMPA receptor content in the postsynaptic density [5]. Our plasticity model (Fig. 1B) describes the temporal dynamics of the weight wjk of a synaptic contact by the dierential equation (1) wjk (t) = F (Sj , Spost, zjk ) − αwjk (t), t d d where F is a functional of the pre- and postsynaptic spike trains Sj and Spost, respectively, and of stochastic synaptic transmission zjk at the contact. The parameter α > 0 describes a slow decay of the weight. The model (1) is a local, spike-timing dependent plasticity rule that we imagine to be realized by the biophysics of dendritic spines in combination with that of the presynaptic terminal. In our model pre- or postsynaptic spikes each leave an exponentially decaying trace (rjk and rpost, respectively) at the synaptic contact (dendritic spine), with a short time constant τ (Fig. 1C). A third lter trace Cjk with a longer time constant τslow tracks the product of rjk · rpost as a measure of the correlation of pre- and postsynaptic spike trains. A fourth lter trace Rpost (with time constant τslow) tracks the postsynaptic ring rate. The combination of these traces gives rise to the STDP model summarized by F in Eq. (1), see Methods for Results 2 Figure 1: Model overview. A: A postsynaptic neuron (post) receives synaptic connections from several excitatory neurons (pre). Each connection consists of multiple synaptic contacts. The synaptic ecacy (weight) of contact k in connection j is denoted as wjk . The sum of contact weights wjk is the total weight wj of this synaptic connection. B: Individual contact weights wjk take continuous, positive values, and change in time according to spike-timing dependent plasticity. A small weight wjk corresponds to a dendritic spine with a small volume, or a thin spine, and a large weight wjk corresponds to a large, or mushroom-shaped dendritic spine. If wjk is greater than 0 we call the contact an actual contact. In contrast, if at any time wjk becomes 0, the contact is pruned and its weight is kept xed at wjk = 0. A contact with wjk = 0 is called a potential (but inactive) contact. It may be transformed into an actual contact by setting wjk to a positive value at random times, with a rate λc (creation rate). C: Components of the plasticity model (top to bottom): presynaptic spike train Sj ; transmitted spike train Sjk at the contact (black, random synaptic failures occur, indicated by black boxes), and its ltered trace rjk (green); postsynaptic spike train Spost (black) and its ltered trace rpost (red); product (correlation) term rjk · rpost composed of pre- and postsynaptic trace (blue); low-pass ltered trace Cjk of the product term rjk·rpost. Synaptic contact plasticity is determined by Cjk . D: Reference distributions of the number of actual (red) [27] and potential (blue) [28] synaptic contacts for pairs of neurons in the adult somatosensory cortex (recurrent connections of layer 5 pyramidal neurons, truncated to nj ≤ 10 and renormalized). The steady state distribution generated by our model is shown in green (data pooled over 150 days of simulation); the distribution of potential contacts in the model is matched to the blue line. 3 further details. Mathematical analysis shows that this plasticity rule establishes a target level for both ring rate and spike-timing correlations (see Model analysis in Methods and Supplementary Information). As a result of ongoing STDP, synaptic model weights can increase or decrease. As soon as a contact weight wjk (t) decreases to zero in the simulation, its dynamics cease and it turns into a potential but inactive contact. However, with a rate of λc = 0.019/day, each potential contact (wjk = 0) may randomly be transformed into an active contact. In such an event the weight wjk of the newly created contact is set to a low, non-zero value wc. After creation the contact weight remains at the value wc for the duration of the period of grace [41] of τgp = 15 min, after which its dynamics again follow (1) (see Fig. 1B). In our model, the number of potential synaptic contacts nj from a presynaptic neuron j to the post- synaptic model neuron lies between 1 to 10, and is distributed as estimated for recurrent connections of layer 5 pyramidal neurons in rat somatosensory cortex [28], so that the blue line in Fig. 1D represents the distribution of potential contacts nj in both model and experiment. The model is insensitive to the exact maximum value of nj , and values of nj up to 20 [28] would give equivalent results, albeit at increased computational cost. In the following, we describe how our plasticity model turns potential contacts into stable weights through the mechanism of STDP-mediated cooperation. Steady state synapse dynamics under STDP-mediated cooperation After an initial transient at the beginning of the simulation, the synaptic contacts uctuate around a steady state. Sample trajectories of all contacts from a single presynaptic neuron to the postsynaptic cell are shown in Fig. 2A-B. All initial contacts of this presynaptic neuron remain over the course of 150 days of simulation, occasionally new contacts are formed, which mature or are quickly removed (see Fig. 2B). The ring rate of the postsynaptic neuron, as well as the total synaptic weight and the average number of contacts are tightly controlled by the plasticity rule (Fig. 2C and Methods). The average number of active contacts hNcontactsi, corresponding to the spine density, is stationary in the model consistent with experiments [42, 12, 43, 5]. Within one day, the relative change of the contact weight ranges from −100% to 500% (Fig. 2D) but the relative uctuations decrease with increasing contact weight, consistent with long-term time-lapse imaging data of dendritic spine volume in vitro [4]. We quantify the statistics of contact weight changes by computing the mean and standard deviation of the change within one day, as a function of the weight. Consistent with experimental measurements of dendritic spine volume [4] the mean contact change (Fig. 2E, black) is positive for an intermediate range of contact weights (spine volumes) and becomes negative for large weights (large volumes), and the standard deviation of the changes (Fig. 2E, red) is rather homogeneous for all weights (volumes) but increases slightly for very large weights. The mean change of contacts of very small weight is negative in our model, in contrast to previous experiments [4], but this dierence might be due to experimental diculties of observing very small spines (see Discussion). As synaptic connections in our model consist of several synaptic contacts, we asked whether there is a systematic relation between synaptic weight and the number of contacts in the steady state (Fig. 2F). Indeed the synaptic weight is strongly correlated with the number of active contacts: Strong synaptic connections consist of at least ve synaptic contacts, and connections with less than three contacts are physiologically weak (Fig. 2F, red, left axis). Moreover, individual contacts in connections made of more than 3 active contacts tend to be stronger than those made of less than 3 (Fig. 2F, blue, right axis). We further assess the statistical properties of the steady state by multiple measures. First, the distribution of active synaptic contact numbers per connection (Fig. 3A) is bimodal, in line with experimental ndings [27, 28], cf. Fig. 1D. Second, the turnover ratio of synaptic contacts in the model is 0.176 ± 0.018/day (mean ± STD) consistent with the values found experimentally in somatosensory cortex [12, 5]. Third, a stable distribution of contacts weights (Fig. 3B) is formed. The probability of a synaptic contact to survive for 8 consecutive days, irrespective of whether the contact is newly created, weak or strong, depends strongly on the number of active contacts in the connection that the contact is part of (Fig. 3C). In other words, contacts within a connection cooperate and stabilize each other. Tracking of individual synaptic contacts that existed at t = 0 (Fig. 3D) reveals a time course of the number of surviving synapses that is qualitatively consistent with long-term in-vivo imaging data of dendritic spines from mouse neocortex [12, 44, 43]. In particular, connections that consist of several contacts are stable for long periods of time (here 150 days) (Fig. 3E inset). The pruned connections are almost exclusively connections that consist of less than four contacts and have relatively small total weight (Fig. 3E). Finally, 4 Figure 2: Dynamics of synaptic contacts in the steady state. A: Example synaptic connection number j = 7, colored lines each correspond to a contact weight wjk over time. New contacts (lled circles) are created with a weight given by the lower dashed line. Long-term stable contacts uctuate around the upper dashed line, which is the xed point of w∗/5 predicted by theory (see Supplementary Information). B: Zoom into A at the time of creation and pruning of two transient synaptic contacts. C: Firing rate Rpost of postsynaptic neuron (black), total synaptic input w =PjPk wjk summed over all presynaptic neurons and contacts (red, unit-less) and average number of active contacts per synaptic connection (green) over time, each sampled in intervals of 6 hours. D: Relative changes of synaptic contact ecacy ∆wjk within one day, each dot corresponds to the change of one contact in one day, data pooled over all contacts and 150 days of simulation (sampled in intervals of 2 days). The horizontal line of dots at ordinate value −1 is due to contacts that were removed within one day. E: Weight dependence of contact changes within one day. Mean µ and standard deviation σ of the change of contact weight ∆wjk within one day, as a function of the contact weight wjk . Both µ and σ were estimated from the data shown in D, grouped into wjk -intervals as indicated (solid lines). Estimates of µ and σ from less than 5 samples were discarded. Error bars denote the standard error of the mean (SEM). For large contact weights σ increases signicantly (∗ indicates p < 0.1 in Welsh's two-sided t-test). F: Total synaptic connection weight wj =Pk wjk (left axis, red) and contact weight wjk (right axis, blue) as a function of the number of actual contacts (wjk > 0), averaged across all synaptic connections. Error bars denote the standard deviation (STD). 5 Figure 3: Statistics of synaptic contacts in the steady state. A: Histogram of the number of active contacts (wjk > 0) of a connection, across all connections j. B: Histogram of contact weights wjk , across all connections j and contacts k. The distribution at day 0 is not signicantly dierent from day 100 (two- sample Kolmogorov-Smirnov test (2sKS), p-value 0.910). C: Fraction of synaptic contacts that survive for 8 days, as a function of the number of actual contacts in the connection, for newly created contacts (gray bars), and existing contacts of weak or strong weight wjk (colors, see legend) in units of 10−3. Survival fractions are estimated separately per day and then averaged over days. Estimates from less than 5 samples are discarded; error bars denote SEM. D: Fraction of surviving actual synaptic contacts that were present at time t = 0 (black: model) in comparison to published experimental dendritic spine survival data in mouse (Hlt '05 S1: somatosensory cortex L5B, age 6 months (exponential decay t) [12]; Hlt '05 VC: visual cortex L5B, age 3-6 months (exponential decay t) [12]; Lws '15 AC: auditory cortex [43]). E: Histogram of the total weights wj of the connections present at t = 0 (black) and of the surviving connections at day 150 (green). Inset: Histogram of the number of contacts of the connections. The removed connections have small total weight and number of contacts; strong connections with many contacts are rarely removed. Note that new contacts created in the meantime are not considered in this analysis. F: Lifetime of entire synaptic connections (dots) during the course of the simulation, as a function of the total connection weight. 6 the lifetime of synaptic connections increases strongly with the total weight of the connection (Fig. 3F), indicating that strong connections are protected against spine turnover by mutual cooperation of contacts. Presynaptic lesions lead to increased formation of persistent contacts Having established that the steady state of our model has characteristics consistent with experimental data, we investigate whether the model can explain experimentally observed transient dynamics of dendritic spines in response to changes of neuronal input. Experimentally, such input changes may be due to trimming whiskers [42, 16], lesions of the retina [44] or occlusion of one of the eyes [45]. Here we model an abstract lesion experiment in which a fraction of the input neurons suddenly cease to re (Fig. 4A). In our lesion model, after ablating 50% of those presynaptic neurons that have active synaptic contacts, we observe a loss of 50% of the synaptic contacts on the postsynaptic neuron after 30 minutes (Fig. 4B). However, this loss is rapidly compensated such that the ring rate and total weight (summed over all presynaptic neurons and all synaptic contacts) are hardly changed throughout the process (Fig. 4B,C). The compensation occurs on two dierent time scales. First, on the time scale of 10 - 30 minutes, existing synaptic contacts are up-regulated from a pre-lesion value between (3.3 ± 1.3) · 10−3 to a value of (6.6 ± 1.3) · 10−3 measured 30 minutes after the lesion (Fig. 4D). Second, on the slow time scale of 10-30 days after the lesion, the number of presynaptic neurons without postsynaptic contact decreases from 886 pre-lesion to 810 thirty days after lesion while the number of presynaptic neurons with one, two, or three contact points transiently increases (Fig. 4E) suggesting that the plasticity rule 'tests' new connection patterns. Competition between synaptic contact points from dierent presynaptic neurons and simultaneous cooperation of synaptic contact arising from the same presynaptic neurons leads to pruning or strengthening, so that after 99 days the distribution of contact numbers and synaptic weights is again very similar (but not yet completely identical) to the distributions pre-lesion (Fig. 4E). The gradual recovery of the contact numbers is due to an elevated probability of newly formed contacts to survive and become long-term stable compared to the control condition (Fig. 4G). In a simulated lesion experiment where 20 percent of the active contacts are removed (instead of 50 percent in the simulations so far), 14.6 percent of newly created contacts survive for 8 days or more, consistent with experimental results on dendritic spines in the somatosensory cortex after whisker trimming [16] and signicantly above the 7.7 percent of surviving contacts in the control condition (Fig. 4G). Synaptic stability and reliability are due to contact multiplicity To investigate the role of multiple synaptic contacts in our model, we compare its dynamics to the same model restricted to single-contact connections (nj = 1 for all j). In the multi-contact model above, with N = 1000 presynaptic neurons there are 4633 potential synaptic contacts in total because every presynaptic neuron makes 4.633 potential contacts on average (Fig. 1D, blue line). To maintain the total number of potential synaptic contacts in the system with single-contact connections we therefore increase the number of presynaptic neurons to N = 4633, keeping all other parameters the same. Similar to the full model, the single-contact model exhibits steady state dynamics (Fig. 5A) in which postsynaptic ring rate, total weight and synaptic contact number are tightly regulated (data not shown), and the distribution of contact weights is stable over time (Fig. 5B). However, in contrast to the multi-contact model (Fig. 3), in the single-contact model the temporal dynamics of synaptic contact survival do not indicate the presence of a subgroup of stable connections (Fig. 5C), and all connections are prone to random pruning, irrespective of whether they have a small or large synaptic weight (Fig. 5D,E). Thus, multiple synaptic contacts are crucial for the long-term stability of strong synaptic connections, through mutual cooperation of the contacts from the same presynaptic neuron: in multi-contact synaptic connections contacts support each other by exciting the postsynaptic neuron more reliably despite random synaptic transmission failures. A theoretical analysis further explains the crucial role of contact multiplicity in synaptic transmission. We characterize the delity of transmission of a presynaptic spike by the ratio of the (trial-averaged) mean and standard deviation of the evoked postsynaptic potential (PSP), which we call the signal-to-noise ratio (see Methods). In our model, failures of synaptic transmission occur randomly at each synaptic contact, which causes variability of the PSP. If a synaptic connection consists of several contacts, a presynaptic spike is more reliably transmitted than in the case of a single contact (Fig. 5F). If 10 contacts are considered to be 7 Figure 4: Rewiring in response to input lesion. A: Schematic of the simulated lesion experiment. A substantial fraction of the presynaptic neurons that have actual synaptic contacts (wjk > 0) onto the postsynaptic cell are ablated (set to very low ring rate of 0.1/s; each connected neuron is ablated with probability plesion = 0.5; unconnected neurons are unaected) at t = 0 (vertical dashes in B-C). B-C: Firing rate Rpost of postsynaptic neuron (black), total synaptic input w =Pj,k wjk summed over all presynaptic neurons and contacts (red, unit-less) and average number of contacts per synaptic connection (green) over time. (cf. Fig. 2C). After the lesion the average number of contacts (green) quickly drops by about 50% (B), and gradually recovers towards the steady state afterwards (C). D-E: Histograms of contact weights wjk (top) and of the number of active contacts (bottom), across all connections and contacts, before and after the lesion. Half of the actual synaptic contacts are removed within 30 minutes after the lesion (D, bottom, green), while the remaining half of the contacts potentiates (D, top, green). Within 99 days the distributions gradually recover (E). F: Example synaptic connection from presynaptic neuron with index j = 18, each colored line corresponds to a contact weight wjk over time, as in Fig. 2A. G: Fraction of newly created persistent contacts (which survive for at least 8 days as dened in [16]), control case (blue, cf. Fig. 2) versus lesion model (red). Two lesion models are shown (red bars), marked with their respective value of plesion. For comparison, data from mouse whisker trimming experiments [16] is shown (open bars, error bars denote STD over observed cells). A smaller lesion with plesion = 0.2 is in better agreement with the experimental data. 8 Figure 5: Single-contact synaptic connections are not long-term stable and have less delity. A-E: Simulation of the plasticity model with only one potential contact per connection (nj = 1 for all j). A: Example synaptic connection number j = 1392, blue line corresponds to contact weight wj1 over time, as in Fig. 2A. B: Histogram of contact weights wj1, across all connections j. A steady state is maintained for 100 days, just as in the case of multiple potential contacts (cf. Fig. 3B). C: Fraction of surviving actual synaptic contacts (solid line) that were present at time t = 0 in comparison to the multi-contact model (dashed line) of Fig. 3D. D: Histogram of the weights wj = wjk of the connections present at t = 0 (black) and of the surviving connections at day 100 (green). Connections of small and large weights are removed unspecically. E: Lifetime of entire synaptic connections during the course of the simulation, as a function of the total connection weight. F: Theoretical signal to noise ratio (SNR) of the postsynaptic potential (13) in response to a presynaptic spike, in connections with multiple actual contacts and stochastic synaptic failures (probability pf). Dashed line indicates the number of actual contacts for which a SNR of 80% of the maximum is achieved, if 10 contacts is considered to be the maximum. 9 the maximum number of active contacts connecting a pair of neurons due to geometrical constraints of the tissue, about 6 contacts give a signal to noise ratio of 80% of the maximum value achievable (irrespective of the synaptic failure rate pf). Note that this number coincides with the typical peak of the model contact numbers in the steady state (cf. Fig. 3A). Indeed, high reliability and low variability of synaptic transmission in layer 5 neurons of the somatosensory cortex of the rat have been observed in glutamate uncaging experiments in acute slices in vitro, and were attributed to the likely presence of multiple synaptic contacts per connection [46]. Network simulations explain reinnervation of input-deprived cortical barrels We wondered whether the structural plasticity rule discussed above would allows us to make predictions of structural changes in a large recurrent network of excitatory and inhibitory neurons. We use a network architecture (Fig. 6A) inspired by rodent barrel cortex with three strongly connected cortical populations (representing the columns corresponding to dierent barrels) of excitatory neurons (exc 1-3), each one preferentially innervated by a thalamic population (tha) that conveys sensory input from one of three whiskers (whi) by strong connections. In the model, connections between dierent cortical populations and from thalamic populations to non-preferred cortical barrel columns are random and weaker on average (Fig. 6D top left), but all excitatory connections whether strong or weak are subject to the same spike-timing dependent structural plasticity model. A short movement (ick) of the whisker is represented in the model by a small increase in the ring rate of the thalamic neurons which results in turn in a modest increase of the ring rate of the corresponding cortical population (Fig. 6B) riding on top of a spontaneous network activity of about 5Hz (Fig. 6D bottom left). After an initial transient of 7 days of simulated time, we followed the mean connection weights of synapses from tha 3 to exc 3, from tha 2 to exc 3, from exc 3 to exc 3, and from exc 2 to exc 2 during 3 days of simulated time and found no changes (Fig 6C, top), indicating that the average connectivity pattern is globally robust during ongoing activity and random whisker stimulation, despite the fact that the structural plasticity rule is always active. After 10 days of simulated time, the whisker corresponding to barrel 3 is trimmed. Whisker lesion is modeled by (i) absence of whisker icks in tha 3 while stimulation of tha 1 and tha 2 continues; and (ii) an exponential decrease of the ring rate of the corresponding thalamic population with a time constant of 5 min to a new baseline level of 0.1Hz. We found that the spiking activity of the network after the lesion remains asynchronous and irregular (Fig. 6B). Within 3 days after the lesion, the recurrent connectivity of the barrel column 3 has increased (Fig. 6C, center) consistent with a recent experiment [47]. The average weight of connections within barrel column 2 is hardly aected by the lesion, but that within barrel column 3 changes substantially. The lateral connections from excitatory neurons of barrel column 2 to 3 and vice versa increase on average. Similarly, the average connection weight of the non-preferred pathway from tha 2 to exc 3 increases whereas the connections in the preferred pathway disappear after the lesion (Fig. 6C, center). We followed the synaptic changes for a total of 50 days after the lesion. Further changes were smaller in magnitude but indicate that the recurrent network slowly continues to reorganize itself into a new connectivity pattern. In particular, the average connectivity from excitatory neurons in barrel column 3 to those in 2 continue to increase. The increase in the average connection weight of the non-preferred pathway (from tha 2 to exc 3) during the rst three days after the lesion could indicate a small increase of all the existing connections, or a stronger increase of a subset of the feed-forward connections. A careful look at the detailed connectivity pattern in Fig. 6D indicates that the latter is the case. Indeed, a subset of neurons located in the excitatory population corresponding to barrel column 3 has become sensitive to stimulation of whisker 2. We rst used a clustering algorithm to identify this subset of neurons and then reordered, and color-coded, the excitatory neurons in barrel column 3 according to their responsiveness (yellow shade: newly responsive to whisker 1; red shade: newly responsive to whisker 2; green shade: responsiveness unchanged). The relevance of a reorganization into subsets manifests itself in four dierent, but consistent ways: (i) the subset of red-shaded neurons in population exc 3 responds more strongly to stimulation of whisker 2 than the average of neurons in exc 3 and nearly as strongly as neurons in the barrel column 2 even though the duration of the response is a bit shorter (Fig. 6E, bottom); (ii) the same subset of red-shaded neurons in population exc 3 has a larger fraction of strong lateral connections to excitatory neurons in barrel column 2 than other neurons in exc 3; (iii) the same subset of red-shaded neurons in population exc 3 receives a larger fraction of strong 10 Figure 6: Structural plasticity in a thalamo-cortical network. A: Schematic of the model. Thalamic neurons (tha) convey sensory input from whiskers (whi) to the recurrent cortical network. Each tha pop- ulation (squares) projects to one of three cortical barrels (circles) of excitatory (exc) neurons. Cortical inhibitory (inh) neurons connect randomly to all barrels. Exc to exc synapses (dotted arrows) are modeled by the structural plasticity model, all other synapses (solid arrows) are static. Plastic exc connections are initialized as follows: Thin arrows wij = 0; thick arrows wij = w∗ (xed point weight, see Supplementary Information). Inh synapses have a constant weight and no synaptic failures. Tha neurons re as Poisson processes but increase their ring rate transiently if the corresponding whisker is icked, which happens ran- domly with rate 1/s. After 10 days of simulation whisker 3 is trimmed (lesion), modeled as a progressive loss of ring of tha 3 neurons (white cross). B: Spike raster plot around time of lesion (dashed vertical line), colors as in A. If a whisker is icked (whi, triangles), the corresponding tha and exc populations respond. C: Relative changes of average connection weights h∆wiji/hwiji between populations. Changes before (top), around time of lesion (center) and long after (bottom). Strong changes during the rst three days post lesion (center) are followed by a slow restructuring process of exc 3 over the following 47 days (bottom). D (top): Exc synaptic connection weights (greyscale, wij ) just before lesion (left) and 50 days after lesion (right). The comparison of the connection weight matrix before and after lesion shows that loss of tha input to exc 3 caused selective rewiring. Exc 3 neurons (401-600) have been clustered in both graphs according to the inputs they receive at simulation end (see Methods, assignments are indicated by shading). D (bottom): Average spike response of exc 2 (black) and exc 3 (blue) neurons in response to whisker 2 icks, just before trimming of whisker 3 (left) and at simulation end (right) (averaged over 60min of recording). Red and yellow colors denote subgroups of exc 3 identied by clustering. Dashed lines: mean ring rate in recording episode. 11 Figure 7: Restructuring can be predicted from initial responses / eect of bounded weights. A: Spike rate in response to whisker 2 icks for each neuron of exc 3, before lesion (black dots) and 50 days post (colored squares), for the thalamo-cortical network simulation shown in Fig. 6. Neurons are ordered according to their pre-lesion response (black). Rates are estimated by counting spikes in a time window of 25ms after each whisker ick. The 100 exc 3 neurons that initially respond strongest to whi 2 are more likely to increase their response to this whisker through structural plasticity (dashed horizontal lines with error bars show mean response 50 days post ± SEM), and more likely to participate in the cluster that is most strongly innervated by this whisker (red; colors correspond to the cluster assignments of the neurons in Fig. 6D, right). B: Rewiring in response to input lesion (as in Fig. 4) with an upper bound of the contact weight so that contact weights cannot grow stronger than twice the xed point value (wjk ≤ 6.4· 10−3). In this case postsynaptic rate and total weight initially decrease in response to the lesion, and recover on a time scale of about 10 hours as new contacts are formed. Simulation data are averaged over consecutive time windows of 1 hour. C: For comparison, the data of the lesion simulation without upper bound of Fig. 4B,C is replotted as in (B). Here the remaining contacts quickly (within less than 1 hour) compensate for the loss of input by strongly increasing their weights, so that the postsynaptic rate shows no visible change. lateral connections from excitatory neurons in barrel column 2 than from neurons in barrel column 1; (iv) the same subset of red-shaded neurons in population exc 3 receives a larger fraction of strong feed-forward connections from thalamic neurons in group 2 than other neurons in exc 3. Taken together, these four observations suggest that the subset of red-shaded neurons in barrel column 3 has been integrated in the information processing stream of barrel column 2. The same observations can be repeated for the yellow- shaded subgroup of neurons in barrel column 3, except that these neurons have been integrated into barrel column 1. In both cases the integration has been made possible by structural plasticity triggered by the lesion. Our simulation results of point (i) above are consistent with experience-dependent receptive eld plastic- ity found experimentally in pyramidal neurons in mouse somatosensory cortex 3-4 days [42] or 20 days [48] after whisker trimming, where neurons that were part of a deprived barrel became responsive to the rst- order surrounding whisker, in particular the subset of neurons located in the border region to the neighboring barrel [48]. Observations (ii) - (iv) listed above are predictions of our model. Note that in our simulations, both barrel columns exc 1 and exc 2 are rst-order surrounding columns of the deprived column (exc 3) since we have not introduced any distance-dependent connectivity. We emphasize that the parameters of the structural plasticity algorithm are kept xed throughout the simulation, be it before, during, or after the lesion: First the network connectivity was stationary before the lesion, then it changed signicantly during 3 days after the lesion, and nally settled into a new state (Fig. 6C) while structural plasticity has always been 'turned on'. Indeed, individual synaptic contact points continue to grow or disappear even during the phases where the coarse connectivity pattern remains unchanged; cf. Fig. 3. Discussion In this study we linked structural dynamics of synaptic contacts in neuronal networks to spike-timing de- pendent plasticity (STDP). The implicit coupling between synaptic contacts onto the same postsynaptic neuron through backpropagating action potentials is sucient to make synaptic contacts from one presy- naptic neuron compete with contacts of other presynaptic neurons and cooperate with contacts of the same 12 presynaptic neuron. The resulting high-dimensional non-linear dynamical system has a steady state with properties consistent with those of excitatory synapses in sensory cortices in terms of (i) a bimodal distri- bution of contact numbers per synaptic connection; (ii) dierences in lifetime between strong and weak connections; and (iii) turnover of dendritic spines. Our model makes at least eight novel predictions. A) A synaptic contact with a given weight is more stable if it is part of a group of four or ve synaptic contacts arising from the same presynaptic neuron than if it is isolated or part of a group of only two synaptic contacts (Fig. 3C). B) The combination of the known result that strong contacts are more stable than small ones (cf. Fig 2D and [4, 12]) with point A, yields the prediction that connections with larger total synaptic weight (summed over all synaptic contacts) are more stable than small ones (Fig. 3F). This could be measured by correlating the survival time of a synaptic contact with the total EPSP amplitude of the connection. C) We predict a substantial fraction of synaptic connections that have only one active contact (see Fig. 3A). These synaptic contacts, however, are small and quickly removed (see Fig. 2F). Since weight, PSP amplitude, volume, and size are tightly correlated [5], these weak synapses might escape electrophysiological or visual detection with standard methods which could explain the dierences to experimental reports [27, 42], but might be detectable using sub-diraction resolution imaging [49, 40]. D) After a lesion, the number of connections consisting of exactly two synaptic contacts increases transiently. While this might be expected since the process of building a novel synaptic connection with ve contacts has to pass through a transient state with only two contacts, the prediction is that only about a quarter of the two-contact connections actually stabilize to a multi-contact synapse, while the majority disappears again. As a consequence, the number of presynaptic neurons without a connection transiently decreases after a lesion before it increases again to a stable value (Fig. 4D and E). E) After whisker trimming, the subset of neurons of the deprived cortical column which will be integrated in the signal processing of an adjacent whisker will establish stronger incoming and lateral projections to the cortical column in which they become integrated than other neurons in the deprived column (Fig. 6D). F) The subset of those neurons of the deprived cortical column that are integrated into a new whisker processing stream do not have to be physical neighbors but can be identied as those who, before the trimming, had already a stronger response to the adjacent whisker (Fig. 7A). G) The fact that, given a connection, the number of synaptic contacts per pre-post pair in experiments peaks around a nite number (e.g. ve contacts) well below the maximal number of potential contacts suggests a novel principle of synaptic plasticity (learning rule) which normalizes the amount of pre-post correlations. This opens a gateway to a new class of learning rules in unsupervised learning which do not maximize second-order correlations (as done by Oja's rule [50]) but normalize these. While experimentally known homeostasis has focused on the normalization of mean ring rate [51] we suggest that it also is worth while to study a normalization of correlations. H) Our learning rule assumes that synaptic contact plasticity strongly depends on local traces of cor- relations that must be processed in the dendritic spines. Therefore, we expect experimental manipulations of biophysical activity traces, such as the calcium concentration, to crucially inuence the development of dendritic spines. In line with this hypothesis, eects of local calcium manipulations on dendritic lopodia have indeed been observed [52]. A recent study on in Drosophila larvae further emphasizes the role of presynaptic neurotransmission for the maturation of synaptic terminals [53]. We combined a specic choice of an STDP model with a point neuron model, but several extensions are possible. First, the synaptic plasticity rule used in this paper could also produce branch specic synaptic plasticity and steady-state congurations [54], by choosing a more complex neuron model with non-linear dendritic branches. Branch specicity of the spine dynamics could further be increased by using a voltage- triggered rule [55] instead of a pure STDP paradigm. Second, while the specic choice of STDP rule used in this paper shows, at low frequencies, a symmetric pre-post and post-pre learning window for LTP we can extend our formalism to more realistic STDP models [56, 57]. Third, we chose a learning rule where the evolution of the synaptic contact weights is formally unbounded, but it is straightforward to also include an upper bound. In our model, an explicit upper bound is not necessary (but see end of this section) because synaptic depression limits further growth of contact weights as soon as the total weight of a connection becomes strong, or individual synaptic contacts get large (see Eq. 7 in Methods and Fig. 2E) [58]. For 13 strong weights our learning rule eectively turns from a Hebbian rule to an anti-Hebbian rule, in the sense that a further increase in correlations of pre- and postsynaptic ring leads to a shrinkage of synaptic weight. Similar principles might explain why, depending on experimental preparations, variable STDP rules have been reported in experiments [59]. Fourth, whereas for our recurrent network simulation of Fig. 6 we have used leaky integrate-and-re neurons, for the mathematical analysis of the system dynamics we chose to describe the activity of the postsynaptic neuron by a Poisson process with a linearly modulated rate. However, because the activity of the postsynaptic model neuron in our simulations uctuates only in a limited range (see Fig. 2C, black), we expect very similar mathematical results in the case of a non-linear neuron linearized around the operating point. Previous structural plasticity models of lesion-induced rewiring assumed a homeostasis of the postsy- naptic ring rate [37] or spike-timing dependent structural plasticity [35, 38]. Both of these approaches, however, are restricted to structural changes, and do not consider the combined dynamics of continuous weights and discrete structural modications, which underlie the concurrent up-regulation of weights in response to the lesion in our model (Fig. 4B). Moreover, previous studies did not consider the interplay of synaptic competition between connections, and cooperation within connections, because they either fo- cused on a single presynaptic neuron with multiple contacts [36, 39], multiple presynaptic neurons with single contacts each [35, 38, 60], or non-competitive, purely homeostatic synaptic dynamics [37, 61]. Previous non-structural models of lesion-induced synaptic plasticity highlighted the time scale of home- ostasis over hours or days [62]. In our model, lesion-induced changes of synaptic contact weights occur rather quickly. While the average synaptic contact number recovers slowly within about 20 days, within less than 10 minutes after the lesion the contact weights from spared presynaptic neurons are upregulated to compensate for the loss of input (see Figures 4B-D and 6C). In apparent contrast, homeostatic synaptic scaling in response to experimental blocking of postsynaptic ring occurs on the time scale of hours [51] or in response sensory deprivation (likely to correspond to a reduction in presynaptic ring) on the time scale of days [62]. In principle, this discrepancy in the speed of the readjustments of the synaptic weights might be alleviated by low-pass ltering the model's weight changes (1) on a time scale of several hours before applying them to the contacts, but that would make plasticity in general unrealistically slow. Instead we propose to extend the model by a combination of hard bounds [63] and multiplicative interaction of Hebbian and explicit homeostatic processes [62]. With hard bounds set to two times the xed point weight, the stationary distributions and results of Figs. 1-3 remain unchanged while the recovery of the ring rate after a lesion is slower since individual synaptic contacts cannot grow beyond the hard bound (see Fig 7B). A more complete model could combine the hard bounds with the essential features of the model of Toyoizumi et al [62]. More generally, the interactions of our rule with manifold plasticity mechanisms such as short- term plasticity [23], inhibitory plasticity [64], homeostasis [51, 37, 62], or intrinsic excitability [65, 60] pose interesting challenges for future research. Methods Spike trains and failures of synaptic transmission. j , t2 j ), where {t1 The activity of the presynaptic neurons 1 ≤ j ≤ N in our model is described by the spike trains Sj (t) = Pm δ(t−tm j , ...} are the spike times of neuron j. All presynaptic spike trains are generated by independent Poisson processes with a constant ring rate of νpre = 5/s. Not all presynaptic spikes, however, are transmitted at each synaptic contact that connects neuron j with the postsynaptic neuron. Some spikes are not transmitted due to random transmission failures [66]. Therefore the spike train transmitted at the contact k is given by Sjk (t) = Xm δ(t − tm j ) zjk (tm j ) . (2) The spike train Sjk (t) = Sj (t) · zjk (t) diers from the presynaptic spike train Sj through multiplication with independent Bernoulli random variables zjk (t) ∈ {0, 1} that describe the stochastic failures of synaptic transmission. In our model, synaptic failures occur randomly and independently with a probability of pf = 0.5, so zjk (tm j ) is 1 (successful transmission) with probability 1 − pf = 0.5. The postsynaptic neuron, in 14 turn, emits the spike train Spost(t) =Pm δ(t − tm on the postsynaptic neuron model are given below. post), where {t1 post, t2 post, ...} are the spike times. Details Slow and fast traces of pre- and postsynaptic activity. At each synaptic contact, traces of pre- and postsynaptic activity are formed, which are illustrated in Fig. 1C. The variables rjk (t) and rpost(t) describe low-pass lters of the pre- and postsynaptic activities Sjk (t) and Spost(t), dened by the dierential equations (3) (4) rjk (t) = −rjk (t) + Sjk (t) , t rpost(t) = −rpost(t) + Spost(t) , d d τ t d d τ with a time constant τ = 20 ms, as is typical for an excitatory postsynaptic potential (EPSP) or an STDP window function. As an estimate of the correlations of pre- and postsynaptic ring on a slower time-scale, each synaptic contact computes Cjk , as well as a slow trace of the postsynaptic activity Rpost. These traces are dened by the dierential equations (5) (6) Cjk (t) = −Cjk (t) + rjk (t) · rpost(t) t Rpost(t) = −Rpost(t) + Spost(t) d d τslow t d d τslow with a time-constant of τslow = 1 min. The choice of Eqs. (3), (4) and (5) leads to a symmetric STDP window, but the formalism can be extended to more standard STDP with long-term potentiation and depression by using more traces with dierent time constants. Synaptic contact plasticity. Synaptic contacts in our model follow a variant of spike-timing dependent plasticity (STDP), see also Fig. 1. Each contact is described by its ecacy (weight) wjk , which is the (unit-less) amplitude of the excitatory postsynaptic potential that the contact k evokes upon arrival of an action potential at the presynaptic terminal and in the absence of synaptic failure. Synaptic contacts evolve according to a local STDP rule 4 C 2 jk (t) − apost 4 R 4 post(t) − αwjk (t), (7) 4 = 0.07506 · 10−6 s3, apost 4 = 0.02016 · 10−6 s3 and with the parameters acorr α = 2 · 10−6 s−1 (see Supplementary Information S1.2 for details on parameter values). As soon as the dynamics (7) lead to a contact of weight wjk (t) ≤ 0 its weight is set to 0 and the dynamics cease (inactive contacts, see also Fig. 1B). Each potential but inactive contact (weight wjk = 0) may be created again at random times tc according to a Poisson process with rate λc = 0.019/day. In such an event, the weight is set to wjk (tc) = wc. Further details on the choice of wc are described below. As suggested by previous works [41, 35] in our model newly created contacts rst pass through a period of grace of duration τgp = 15 min, during which the weight is xed to wc. After the period of grace has passed (for t ≥ tc + τgp), the weight dynamics again follow Eq. (7). In the event of synaptic contact creation (at tc) the internal state variables rjk (tc), rpost(tc), Cjk (tc) and Rpost(tc) of the contact are each initialized to zero. The period of grace serves as a protected time interval for these variables to equilibrate to the current system state  i.e. to obtain a good estimate of the present pre- and postsynaptic spike rates and correlations. wjk (t) = acorr 2 Cjk (t) − acorr 2 = 1.94569 · 10−6 s, acorr t d d Model of postsynaptic activity. For the simulation of the recurrent network in Fig. 6, the pre- and postsynaptic activity was generated by leaky integrate-and-re neuron models. However, to perform a mathematical analysis of the model's 15 dynamics, we assume a minimal model of the postsynaptic neuron for the remainder of the study. In the absence of synaptic input from the presynaptic neurons, the postsynaptic neuron res with a baseline ring rate λ0 = 1/s (as a Poisson process). We further assume that synaptic inputs cause transient increases of the ring rate on the typical time-scale τ of an EPSP. We thus model the dynamics of the postsynaptic neuron's ring rate λ(t) as N X j=1 njXk=1 wjk (t)Sjk (t − d) , (8) λ(t) = −(λ(t) − λ0) + t d d τ where the second term sums all inputs across all nj synaptic contacts over all N presynaptic neurons, and d = 1ms denotes the synaptic transmission delay. Potential synaptic contacts. Each presynaptic neuron j may be connected to the postsynaptic neuron by several synaptic contacts, up to a maximum of nj , 1 ≤ k ≤ nj , cf. Fig. 1A. The number nj of potential contacts of a connection is random, with a probability distribution (Fig. 1D, blue line) estimated in [28] for synapses connecting Layer 5 pyramidal neurons within a maximum distance of 50 µm in rat barrel cortex. For computational reasons we limited n to a maximum of 10 and renormalized the distribution P(n). Accordingly each value of 1 ≤ n ≤ 10 should appear (in expectation) P(n) · N times, where N is the number of presynaptic neurons. In the model we randomly assigned P(n) · N presynaptic neurons to each potential contact number n in order to exactly reproduce P(n), the reported distribution of contact numbers per connection, except for Fig. 5 (single contact case) and Fig. 6 (recurrent network). Simulation of the plasticity model. We performed simulations of the full system using the Neural Simulation Tool (NEST) [67]. We implemented the model as a new multi-contact synapse object, using analytical integration of Eq. (7) as described in Supplementary Information S1.4. Spikes were restricted to a simulation time grid with a step size of ∆t = 1 ms. The complete system state (wjk , rjk , rpost, Cjk , Rpost, λ) was recorded in intervals of 5 min. The source code of our plasticity model will be made available on-line as part of NEST (as 'stdp_spl_synapse_hom') upon acceptance of this manuscript for publication; referees can get access by request to the journal editor. To be able to compare the model dynamics to the adult networks of the referenced experimental studies, we initially simulated the model until a stable synaptic conguration was reached. This steady-state of the system, which is the state at t = 0 in Figs. 2 to 4, was obtained by simulating for 100 days after initialization at the theoretically determined xed point. The initial state was set to wjk = w∗/5 for 1 ≤ j ≤ 100 and 1 ≤ k ≤ 5, and wjk = 0 else (see details on the xed point and denition of w∗ below). However, because the numbering of presynaptic neurons in the model is random and arbitrary, it is possible that initially not all presynaptic neurons 1 ≤ j ≤ 100 have at least nj ≥ 5 potential contacts. Therefore, for any presynaptic neuron 1 ≤ j ≤ 100 for which nj < 5, we looked for another neuron j0 > 100 with nj0 ≥ 5, and exchanged nj and nj0. This procedure allowed us to initialize the system to the theoretical equilibrium state. We also checked that if the system is initialized unconnected (wjk = 0 for all j, k), fully connected, or randomly connected, the postsynaptic rate, the total weight, and the contact numbers approach a similar steady state (data not shown). (9) wjk (t)i ≈ acorr 2 hCjk (t)i − acorr 4 hCjk (t)i2 − apost 4 hRposti4 − αwjk , t d d h We derive the expected dynamics of the weight of a single synaptic contact under this model. Since dendritic spine plasticity is a slow process compared to the dynamics of action potentials and synaptic transmission, we take the average, denoted as h·i, of Eq. (7) over realizations of the spike trains (S) and synaptic transmission failures (z). We obtain Expected dynamics of synaptic contacts. 16 which is approximate because squaring and averaging of the terms Cjk and Rpost have been interchanged. The terms in Eq. (9) can be evaluated as (see Supplementary Information S1.1 for the derivation) hRposti = hSposti = hλi = λ0 + νpre(1 − pf) hCjki = hrjk rposti = hrjk λi = νpre(1 − pf) · " 1 2τ e−d/τ (pfwjk + (1 − pf) N j=1 wjk njXk=1 X wjl ) + hRposti# . njXl=1 (10) (11) Eq. (10) establishes that the postsynaptic rate is determined by the sum of all synaptic weights. Since we assume pre- and postsynaptic neurons to be of the same type, we have chosen the parameters of the plasticity rule such that this rate is on average hRposti ≈ νpost = νpre. Our simulations show that this value is tightly maintained. This implies that the sum of weightsPjPk wjk is normalized [68], cf. Eq. (10). Indeed previous theoretical work [58] has shown that terms like −R 4 post in our learning rule lead to a normalization of the total weight. According to Eq. (11), for small transmission failure probability pf → 0 the dynamics of the contact wjk (9) is dominated by the total weight wj =Pk wjk . For large pf → 1, the evolution of wjk is independent of wj ; so increasing the failure probability pf gradually decouples the dynamics of the contacts. Inserting Eqs. (10) and (11) into Eq. (9) yields a closed, non-linear system of ordinary dierential dynamical equations in wjk (t), with 1 ≤ k ≤ nj and 1 ≤ j ≤ N. In Supplementary Information we describe how the xed points of the expected contact dynamics can be analyzed and used to calibrate the system (S1.2), and how prototypical trajectories of newly created contacts can be constructed (S1.3). Cooperation and competition. Eqs. (9)-(11) enable us to illustrate the process of cooperation and competition, closely linked to the stabilization of the postsynaptic rate and correlations. The postsynaptic rate does not depend on the weight of any specic synaptic contact, but only on the total input, summed over all weights and contacts; cf. Eq. 10. By contrast, Eq. (11) depends not only on the total input via the rate Rpost, but in addition also on the individual weight wjk and the total weight wj = Pl wjl arising from the same presynaptic neuron. To study competition, let us consider a uniform state and suppose that all correlations Cjk = c and all momentary weights wjk = w for all j, k are small but positive. With α (cid:28) 1 in Eq. (9), the dominant evolution is therefore an increase of all weights, driven by the term acorr c. However, as the weights increase, the ring rate does so as well and therefore the term hRposti4 eventually stops further growth. This is the essential step of ring rate stabilization. For the same ring rate hRposti, some weights will grow further at the expense of others inducing competition via the instability of the uniform state, just as in other models [50]. The instability is caused by a positive feedback loop between hdwjk (t)/dti on the left-hand side of Eq. (9) and wjk on the right-hand side of Eq. (11). Going beyond standard plasticity models, Eq. (11) shows that correlations Cjk driving the contact weight wjk increase not only proportionally to this specic contact, but also increase with the weight of other contactsPl wjl from the same presynaptic neuron. The positive dependence gives rise to cooperation between contacts arising from the same neuron. The optimal hCjki2 in Eq. (9). The amount of correlation, and hence cooperation, however, is limited by the term −acorr interplay of Eqs. (9)-(11) therefore stabilizes the ring rate, or total input PjPk wjk , as well as the total amount of correlations in active contacts, or total weight Pk wjk , of those presynaptic neurons that have at least one active contact. 2 4 Contact creation. With a rate of λc each potential but inactive contact (wjk = 0) may be randomly transformed into an active contact. In such an event, called creation, its weight wjk is set to a low, non-zero value wc, and after the period of grace has elapsed, its dynamics follow Eq. (7) (see Fig. 1B). We have adjusted the value of λc to t the experimentally measured turnover ratio of dendritic spines as follows. 17 The turnover ratio, dened as TOR = (#created + #removed)/(2 · #total · day), was found to be 0.154/day in the adult somatosensory cortex [12], where #created, #removed and #total are the numbers of spines that were created, removed, and the total number observed in one day, respectively. In a steady state we additionally expect that #created = #removed, which implies TOR = #created/(#total · day). Now assume that the model is in a steady-state with a distribution of actual contacts as in Fig. 1D, red line. We derive the creation rate λc loosely from this distribution, by assuming that approximately 10% of synaptic connections have active contacts, and these have about 5 active contacts. Counting the active synaptic contacts in this case, one would observe about #total ≈ 10% · N · 5 = 0.5 · N. On the other hand, the total number of potential contacts in our model is #potential ≈ 4.6 · N (Fig. 1D, blue, shows the histogram of nj across connections). Thus, the number of inactive contacts that are available for creation is #potential − #total. Hence, the expected number of creations is #created = (#potential − #total) · λc · 1 day = (4.6 − 0.5) · N · λc · 1 day. Inserting #created and #total into the expression for TOR above, and solving for λc, we obtain λc = TOR · 0.5/4.1 = 0.019/day. To better understand the eects of multiple synaptic contacts in presence of stochastic transmission, let us analyze the postsynaptic response. To this aim, we force the presynaptic neuron j to emit an additional spike at time tpre. For convenience we neglect the synaptic transmission delay here (d → 0), which has no eect on the following reasoning. Assuming constant weights on the short time-scale of synaptic signaling τ , and by averaging Eq. (S2) over all other presynaptic spikes and their synaptic transmission, we obtain the transient postsynaptic response wjk zjk (tpre), (12) njXk=1 θ(t − tpre)e−(t−tpre)/τ 1 τ Lj (t tpre) = νpost + where we also inserted hRposti ≈ νpost. Note that Lj (t tpre) is still a stochastic quantity due to stochastic synaptic transmission zjk (tpre) at the contacts k. We may obtain the mean spike-triggered response by averaging over the remaining stochasticity of zjk Signal-to-noise ratio of synaptic responses. θ(t − tpre)e−(t−tpre)/τ (1 − pf)wj . 1 τ hLj (t tpre)i = νpost + Thus the average response only depends on the total synaptic weight wj , and not on the conguration of contact weights wjk . Similarly, the variance of the response can be derived as var[Lj (t tpre)] = pf(1 − pf) 1 τ 2 θ(t − tpre)e−2(t−tpre)/τ njXk=1 w 2 jk . Here we see that the contact conguration wjk determines the variance of the postsynaptic response. To further understand these properties, consider a synaptic weight wj that is made of nj contacts of weight j /nj . For this jk = w 2 k=1 w 2 case we evaluate the signal-to-noise ratio of the synaptic response as wjk = wj /nj . Then the sum of squared weights term in var[Lj (t)] becomes Pnj · nj . =s 1 − pf pf SNRj = hLj (t tpre)i − νpost qvar[Lj (t tpre)] (13) Therefore, in presence of synaptic transmission failures, multiple synaptic contacts increase the signal-to- noise ratio of synaptic transmission, proportional to the square root of the number of contacts. Previously, a related result has been found numerically for the mutual information of synaptic inputs and neural outputs via multiple contacts [69]. Recurrent network model Here we describe the thalamo-cortical network model presented in Fig. 6. The recurrent cortical network consists of NB = 3 barrels of NE = 200 excitatory (exc) neurons each, and NI = 200 inhibitory (inh) 18 neurons that connect without preference to all the barrels (random connections, see details below). All cortical neurons are modeled as leaky integrate-and-re (LIF) neurons with alpha-function shaped postsy- naptic currents ('iaf_psc_alpha' neuron model in NEST simulator). The parameters of the LIF neurons are: membrane time constant τLIF = 20.6ms, reset and resting potential −70mV, action potential threshold −55mV, synaptic time constant 2ms, refractory period 2ms. There are NB · NE + NI = 800 cortical neurons in total. All synaptic delays are 1ms. Each barrel of excitatory neurons is further innervated by thalamic inputs that convey information from the whiskers. There are NT = 100 thalamic input neurons (tha) per barrel. All NB · NT = 300 tha neurons are modeled as excitatory linear Poisson neurons according to Eq. 8, with a baseline ring rate of λ0 = 4.5/s. Each group of thalamic neurons modulates its ring rate in response to icks of the corresponding whiskers (whi). The sequence of whisker icks is stochastic and described by Poisson processes S1, S2, S3 with rate νwhi = 1/s. The connection weights from whi to tha are chosen as wwhi = 0.5. We assume full connectivity to the respective tha populations, such that each tha neuron responds to each ick of the corresponding whisker. In response to a whisker ick, tha neurons of the receiving population increase their ring rate λ(t) transiently by wwhi/τ = 25/s, and subsequently their rate decays back to λ0 quickly with time constant τ (cf. Eq. (8)). The structural plasticity model (7) describes all exc to exc connections, both thalamo-cortical and intra- cortical, and is continuously active (except for the rst hour after simulation); autapses are excluded. All connections (both from tha to exc and exc to exc) have potential synaptic contacts, but initially the network is connected in a whisker-specic manner as depicted in Fig. 6A, see also below. Because the recurrent network contains many postsynaptic neurons, we need two indices to name a synaptic connection. A contact weight here is denoted by wijk instead of wjk above, where i denotes the postsynaptic and j denotes the presynaptic neuron, and k the contact. Accordingly, the plasticity rule (7) here reads (14) wijk (t) = acorr 2 Cijk (t) − acorr 4 C 2 ijk (t) − apost 4 R 4 i (t) − αwijk (t), t d d and the total weight of a synaptic connection is wij (t) =Pnij k=1 wijk (t), where nij is the number of potential contacts for the connection from j to i. For simplicity all nij are drawn from the probability distribution P(nij ) (Fig. 1D, blue), irrespective of which group (exc or tha) the neurons i and j belong to. Because the postsynaptic neurons here are LIF neurons, synaptic ecacies wijk have to be expressed in units of the PSP (in contrast, above wjk is a unit-less quantity). To match the impulse response function of the LIF neurons receiving an input spike with the xed point weight w∗ to the response of the linear Poisson neurons used above, we scale the synaptic weights as wijk = γ · wijk , with γ = 62.82 mV, leading to a typical EPSP amplitude of γw∗ = 1.01 mV. Substituting wijk into Eq. (14) implies that, to maintain the same plasticity dynamics as above, the parameters of the learning rule have to be rescaled according to acorr , w0 7→ γw0. We further inject additional Poisson excitatory and inhibitory input spikes to all LIF neurons, with synaptic weights γw∗ (excitation) and −4γw∗ (inhibition). Exc neurons receive input rates 1519.2/s (excitation) and 328.3/s (inhibition), inh neurons receive 1391.0/s (excitation) and 351.2/s (inhibition). The scaling factor γ and the Poisson process input rates were numerically optimized to match the dynamics of the LIF neuron model to that of the linear Poisson neuron model used above. All other parameters take the same values as before. Note that the network parameters here are chosen such that the system operates approximately at the xed point of the plasticity dynamics analyzed in Fig. S1. 7→ γapost 7→ γacorr 7→ γacorr , apost 4 , acorr 4 2 2 4 4 Apart from its (NB[NT + NE])(NB[NE − 1]) ≈ 5.4 · 105 plastic excitatory connections our network also has static synapses. These we set as follows. We choose a connection probability of pconn = 1/3. Each excitatory neuron receives pconnNI synapses from randomly chosen inh neurons with a xed weight of −(1 − pf)γgw∗, with g = 2.5. Each inhibitory neuron also receives this amount of inhibitory synapses, and pconnNE excitatory synapses from randomly chosen exc neurons from each of the NB cortical barrels, with a xed weight of (1 − pf)γw∗ (these synapses have no transmission failures, therefore the weight is scaled down by the expected transmission rate (1 − pf) of the plastic, stochastic ones). We initialize the plastic synapses at the theoretically derived xed point w∗, with an expected total of 100 active input connections with 5 contacts each per neuron. So, for each exc-exc and tha-exc connection, we set wij (0) = γw∗qij if neuron i and neuron j belong to the same barrel, where qij is a Bernoulli random number that is 1 with probability pconn and 0 else (in this way, we get (NT + NE)pconn = 100 incoming 19 connections per exc neuron in expectation). If i and j are part of dierent barrels, we set wij (0) = 0. If there are ve or more potential contacts (nij ≥ 5) in connection i, j, we set wijk (0) = wij (0)/5 for 1 ≤ k ≤ 5 and wijk (0) = 0 for k ≥ 5. If there are less contacts (nij < 5) but wij (0) > 0, we set wij (0) = 0 and look for a connection i0, j0 that connects the same two groups, has wi0j0(0) = 0 and (ni0j0 ≥ 5), and we set wi0j0(0) = γw∗ for this connection instead. In Fig. 6D neurons of barrel column exc 3 are ordered according to labels obtained by clustering. We used feature agglomeration based on Ward's hierarchical clustering [70] to assign one of three cluster labels to the vector {wi1, wi2, ...} of connection weights (at simulation end) from any exc neuron onto each exc 3 neuron i. Simulations were performed using NEST [67] as described above, except for: (i) contact weights were not recorded, merely the total connection weights wij (t) and the number of active contacts of each connection; (ii) the system state was recorded only every 60min instead of every 5min above; (iii) we simulated using 44 computing cores in parallel instead of a single core. Acknowledgments: Research was supported by the European Union Seventh Framework Program (FP7) under grant agreement no. 604102 (Human Brain Project, M.D.) and by the Swiss National Science Foundation (200020_147200, A.S.). References [1] Sheng M, Hoogenraad CC (2007) The postsynaptic architecture of excitatory synapses: a more quan- titative view. Annu Rev Biochem 76: 823847. [2] Yuste R (2011) Dendritic spines and distributed circuits. Neuron 71: 772781. [3] Bonhoeer T, Yuste R (2002) Spine motility. phenomenology, mechanisms, and function. Neuron 35: 10191027. [4] Yasumatsu N, Matsuzaki M, Miyazaki T, Noguchi J, Kasai H (2008) Principles of long-term dynamics of dendritic spines. J Neurosci 28: 1359213608. [5] Holtmaat A, Svoboda K (2009) Experience-dependent structural synaptic plasticity in the mammalian brain. Nat Rev Neurosci 10: 647658. [6] Kasai H, Fukuda M, Watanabe S, Hayashi-Takagi A, Noguchi J (2010) Structural dynamics of dendritic spines in memory and cognition. Trends Neurosci 33: 121129. [7] Loewenstein Y, Kuras A, Rumpel S (2011) Multiplicative dynamics underlie the emergence of the log-normal distribution of spine sizes in the neocortex in vivo. J Neurosci 31: 94819488. [8] Sala C, Segal M (2014) Dendritic spines: the locus of structural and functional plasticity. Physiol Rev 94: 141188. [9] Kalisman N, Silberberg G, Markram H (2005) The neocortical microcircuit as a tabula rasa. Proc Natl Acad Sci U S A 102: 880885. [10] Stepanyants A, Hof PR, Chklovskii DB (2002) Geometry and structural plasticity of synaptic connec- tivity. Neuron 34: 275288. [11] Kappel D, Habenschuss S, Legenstein R, Maass W (2015) Network plasticity as bayesian inference. PLoS Comput Biol 11: e1004485. [12] Holtmaat AJGD, Trachtenberg JT, Wilbrecht L, Shepherd GM, Zhang X, et al. (2005) Transient and persistent dendritic spines in the neocortex in vivo. Neuron 45: 279291. 20 [13] Grutzendler J, Kasthuri N, Gan WB (2002) Long-term dendritic spine stability in the adult cortex. Nature 420: 812816. [14] Hayashi-Takagi A, Yagishita S, Nakamura M, Shirai F, Wu YI, et al. (2015) Labelling and optical erasure of synaptic memory traces in the motor cortex. Nature 525: 333338. [15] Matsuzaki M, Honkura N, Ellis-Davies GCR, Kasai H (2004) Structural basis of long-term potentiation in single dendritic spines. Nature 429: 761766. [16] Holtmaat A, Wilbrecht L, Knott GW, Welker E, Svoboda K (2006) Experience-dependent and cell- type-specic spine growth in the neocortex. Nature 441: 979983. [17] Zito K, Scheuss V, Knott G, Hill T, Svoboda K (2009) Rapid functional maturation of nascent dendritic spines. Neuron 61: 247258. [18] Kwon HB, Sabatini BL (2011) Glutamate induces de novo growth of functional spines in developing cortex. Nature 474: 100104. [19] Minerbi A, Kahana R, Goldfeld L, Kaufman M, Marom S, et al. (2009) Long-term relationships between synaptic tenacity, synaptic remodeling, and network activity. PLoS Biol 7: e1000136. [20] Fusi S, Drew PJ, Abbott LF (2005) Cascade models of synaptically stored memories. Neuron 45: 599611. [21] Clopath C, Ziegler L, Vasilaki E, Büsing L, Gerstner W (2008) Tag-trigger-consolidation: a model of early and late long-term-potentiation and depression. PLoS Comput Biol 4: e1000248. [22] Roxin A, Fusi S (2013) Ecient partitioning of memory systems and its importance for memory con- solidation. PLoS Comput Biol 9: e1003146. [23] Zenke F, Agnes EJ, Gerstner W (2015) Diverse synaptic plasticity mechanisms orchestrated to form and retrieve memories in spiking neural networks. Nat Commun 6: 6922. [24] Kempter R, Gerstner W, van Hemmen JL (1999) Hebbian learning and spiking neurons. Phys Rev E 59. [25] Abbott LF, Nelson SB (2000) Synaptic plasticity: taming the beast. Nat Neurosci 3: 11781183. [26] van Rossum MC, Bi GQ, Turrigiano GG (2000) Stable hebbian learning from spike timing-dependent plasticity. J Neurosci 20: 88128821. [27] Markram H, Lübke J, Frotscher M, Roth A, Sakmann B (1997) Physiology and anatomy of synaptic connections between thick tufted pyramidal neurones in the developing rat neocortex. J Physiol 500 ( Pt 2): 409440. [28] Fares T, Stepanyants A (2009) Cooperative synapse formation in the neocortex. Proc Natl Acad Sci U S A 106: 1646316468. [29] Markram H, Muller E, Ramaswamy S, Reimann MW, Abdellah M, et al. (2015) Reconstruction and simulation of neocortical microcircuitry. Cell 163: 456492. [30] Reimann MW, King JG, Muller EB, Ramaswamy S, Markram H (2015) An algorithm to predict the connectome of neural microcircuits. Front Comput Neurosci 9: 120. [31] Stuart G, Schiller J, Sakmann B (1997) Action potential initiation and propagation in rat neocortical pyramidal neurons. J Physiol 505 ( Pt 3): 617632. [32] Markram H, Lübke J, Frotscher M, Sakmann B (1997) Regulation of synaptic ecacy by coincidence of postsynaptic aps and epsps. Science 275: 213215. 21 [33] Petersen CC, Malenka RC, Nicoll RA, Hopeld JJ (1998) All-or-none potentiation at ca3-ca1 synapses. Proc Natl Acad Sci U S A 95: 47324737. [34] O'Connor DH, Wittenberg GM, Wang SSH (2007) Timing and contributions of pre-synaptic and post- synaptic parameter changes during unitary plasticity events at ca3-ca1 synapses. Synapse 61: 664678. [35] Helias M, Rotter S, Gewaltig MO, Diesmann M (2008) Structural plasticity controlled by calcium based correlation detection. Front Comput Neurosci 2: 7. [36] Deger M, Helias M, Rotter S, Diesmann M (2012) Spike-timing dependence of structural plasticity explains cooperative synapse formation in the neocortex. PLoS Comput Biol 8: e1002689. [37] Butz M, van Ooyen A (2013) A simple rule for dendritic spine and axonal bouton formation can account for cortical reorganization after focal retinal lesions. PLoS Comput Biol 9: e1003259. [38] Vlachos A, Helias M, Becker D, Diesmann M, Deller T (2013) Nmda-receptor inhibition increases spine stability of denervated mouse dentate granule cells and accelerates spine density recovery following entorhinal denervation in vitro. Neurobiol Dis 59: 267276. [39] Fauth M, Wörgötter F, Tetzla C (2015) The formation of multi-synaptic connections by the inter- action of synaptic and structural plasticity and their functional consequences. PLoS Comput Biol 11: e1004031. [40] Attardo A, Fitzgerald JE, Schnitzer MJ (2015) Impermanence of dendritic spines in live adult ca1 hippocampus. Nature 523: 592596. [41] Le Bé JV, Markram H (2006) Spontaneous and evoked synaptic rewiring in the neonatal neocortex. Proc Natl Acad Sci U S A 103: 1321413219. [42] Trachtenberg JT, Chen BE, Knott GW, Feng G, Sanes JR, et al. (2002) Long-term in vivo imaging of experience-dependent synaptic plasticity in adult cortex. Nature 420: 788794. [43] Loewenstein Y, Yanover U, Rumpel S (2015) Predicting the dynamics of network connectivity in the neocortex. J Neurosci 35: 1253512544. [44] Keck T, Mrsic-Flogel TD, Vaz Afonso M, Eysel UT, Bonhoeer T, et al. (2008) Massive restructuring of neuronal circuits during functional reorganization of adult visual cortex. Nat Neurosci 11: 11621167. [45] Rose T, Jaepel J, Hübener M, Bonhoeer T (2016) Cell-specic restoration of stimulus preference after monocular deprivation in the visual cortex. Science 352: 13191322. [46] Nawrot MP, Schnepel P, Aertsen A, Boucsein C (2009) Precisely timed signal transmission in neocortical networks with reliable intermediate-range projections. Front Neural Circuits 3: 1. [47] Albieri G, Barnes SJ, de Celis Alonso B, Cheetham CEJ, Edwards CE, et al. (2015) Rapid bidirectional reorganization of cortical microcircuits. Cereb Cortex 25: 30253035. [48] Wilbrecht L, Holtmaat A, Wright N, Fox K, Svoboda K (2010) Structural plasticity underlies experience- dependent functional plasticity of cortical circuits. J Neurosci 30: 49274932. [49] Nägerl UV, Willig KI, Hein B, Hell SW, Bonhoeer T (2008) Live-cell imaging of dendritic spines by sted microscopy. Proc Natl Acad Sci U S A 105: 1898218987. [50] Oja E (1982) A simplied neuron model as a principal component analyzer. J Math Biol 15: 267273. [51] Ibata K, Sun Q, Turrigiano GG (2008) Rapid synaptic scaling induced by changes in postsynaptic ring. Neuron 57: 819826. [52] Lohmann C, Finski A, Bonhoeer T (2005) Local calcium transients regulate the spontaneous motility of dendritic lopodia. Nat Neurosci 8: 305312. 22 [53] Choi BJ, Imlach WL, Jiao W, Wolfram V, Wu Y, et al. (2014) Miniature neurotransmission regulates drosophila synaptic structural maturation. Neuron 82: 618634. [54] Druckmann S, Feng L, Lee B, Yook C, Zhao T, et al. (2014) Structured synaptic connectivity between hippocampal regions. Neuron 81: 629640. [55] Clopath C, Büsing L, Vasilaki E, Gerstner W (2010) Connectivity reects coding: a model of voltage- based stdp with homeostasis. Nat Neurosci 13: 344352. [56] Pster JP, Gerstner W (2006) Triplets of spikes in a model of spike timing-dependent plasticity. J Neurosci 26: 96739682. [57] Senn W, Markram H, Tsodyks M (2001) An algorithm for modifying neurotransmitter release probability based on pre- and postsynaptic spike timing. Neural Comput 13: 3567. [58] Zenke F, Hennequin G, Gerstner W (2013) Synaptic plasticity in neural networks needs homeostasis with a fast rate detector. PLoS Comput Biol 9: e1003330. [59] Sjöström PJ, Turrigiano GG, Nelson SB (2001) Rate, timing, and cooperativity jointly determine cortical synaptic plasticity. Neuron 32: 11491164. [60] Miner D, Triesch J (2016) Plasticity-driven self-organization under topological constraints accounts for non-random features of cortical synaptic wiring. PLoS Comput Biol 12: e1004759. [61] Diaz-Pier S, Naveau M, Butz-Ostendorf M, Morrison A (2016) Automatic generation of connectivity for large-scale neuronal network models through structural plasticity. Front Neuroanat 10: 57. [62] Toyoizumi T, Kaneko M, Stryker MP, Miller KD (2014) Modeling the dynamic interaction of hebbian and homeostatic plasticity. Neuron 84: 497510. [63] Morrison A, Diesmann M, Gerstner W (2008) Phenomenological models of synaptic plasticity based on spike timing. Biol Cybern 98: 459478. [64] Vogels TP, Sprekeler H, Zenke F, Clopath C, Gerstner W (2011) Inhibitory plasticity balances excitation and inhibition in sensory pathways and memory networks. Science 334: 15691573. [65] Daoudal G, Debanne D (2003) Long-term plasticity of intrinsic excitability: learning rules and mecha- nisms. Learn Mem 10: 456465. [66] Hessler NA, Shirke AM, Malinow R (1993) The probability of transmitter release at a mammalian central synapse. Nature 366: 569572. [67] Bos H, Morrison A, Peyser A, Hahne J, Helias M, et al. (2015). Nest 2.10.0. doi:10.5281/zenodo.44222. URL http://dx.doi.org/10.5281/zenodo.44222. [68] Kempter R, Gerstner W, van Hemmen JL (2001) Intrinsic stabilization of output rates by spike-based hebbian learning. Neural Comput 13: 27092741. [69] Zador A (1998) Impact of synaptic unreliability on the information transmitted by spiking neurons. J Neurophysiol 79: 12191229. [70] Pedregosa F, Varoquaux G, Gramfort A, Michel V, Thirion B, et al. (2011) Scikit-learn: Machine learning in Python. Journal of Machine Learning Research 12: 28252830. 23 Supplementary Information for Multi-contact synapses for stable networks: a spike-timing dependent model of dendritic spine plasticity and turnover Moritz Deger1,2, Alexander Seeholzer1, Wulfram Gerstner1 1: School of Computer and Communication Sciences and School of Life Sciences, Brain Mind Institute, École Polytechnique Fédérale de Lausanne, 1015 Lausanne EPFL, Switzerland 2: Institute for Zoology, Faculty of Mathematics and Natural Sciences, University of Cologne, 50674 Cologne, Germany To derive Eqs. (10) and (11) in the main text, we rst write the solution of the dierential equations of the traces dened in Eqs. (3) and (4) of the main text, (S1) e−(t−s)/τ S·(s)ds , t -∞ 1 τ r·(t) = where · stands for either jk or post. So both rjk and rpost are low-pass ltered versions of presynaptic (Sjk ) and postsynaptic activity (Spost), respectively, with time constant τ . Similarly, the expressions Cjk (t) and Rpost(t) in Eq. (7) of the main text denote low-pass ltered traces (with time-constant τslow) of pre-post correlations and the postsynaptic spike train, respectively. Because presynaptic ring and synaptic failures are modeled as independent random variables, taking the expectation value of rjk (S1) with respect to the realizations of Sjk yields hrjki = hSjki = νpre(1 − pf). For simplicity, we have assumed that the traces r·(t) as well as the postsynaptic rate λ(t) all evolve on the common time scale τ . The postsynaptic activity λ(t) can then, by solving Eq. (8) of the main text and inserting (S1), be written as S1 Supplementary Methods S1.1 Derivation of pre-post correlations. wjk (t)rjk (t − d), (S2) e−(t−s)/τ N X j=1 njXk=1 wjk (s)Sjk (s − d)ds t -∞ njXk=1 X 1 τ j=1 N λ(t) = λ0 + = λ0 + where in the second step we assumed that wjk does not change much on the (fast) time-scale of τ . Since the expectation of rjk is νpre(1 − pf) we obtain Eq. (10) of the main text. We arrive at the expression hrjk λi, which evaluates to Analogously, we compute Eq. (11) of the main text by inserting the explicit solutions of the traces (S1). ds e−(t−s)/τ 1 τ t -∞ du e−(t−u)/τXi,l wilhSjk (s)Sil (u − d)i + hSjk (s)iλ0 . (S3) t -∞ 1 τ hrjk λi = Here we insert hSjk (s)Sil (u)i = hSj (s)zjk (s)Si (u)zil (u)i = hSj (s)Si (u)ihzjk (s)zil (u)i = Cji (s, u)[δkl (1 − pf) + (1 − δkl )(1 − pf)2] + ν2 pre(1 − pf)2 , 1 where we substituted the covariance function of the spike trains j and i, Cji (s, u) = hSj (s)Si (u)i − hSj (s)ihSi (u)i. For the Poisson processes of our model we have Cji (s, u) = νpreδij δ(s − u). Here δij denotes the Kronecker symbol and δ(·) denotes the Dirac delta function. When summed over i, l together with wil , the expression above becomes X i,l wilhSjk (s)Sil (u)i =Xi,l wilCji (s, u)[δkl (1 − pf) + (1 − δkl )(1 − pf)2] + ν2 pre(1 − pf)2Xi [pfwik + (1 − pf)wi ]Cji (s, u) + νpre[hRposti − λ0]! = (1 − pf) Xi wi (S4) (S5) = νpre(1 − pf) [δ(s − u) (pfwjk + (1 − pf)wj ) + hRposti − λ0] , where we inserted the covariance function Cji for independent Poisson process inputs in the last step. By insertion of (S5) into (S3) the last equality of Eq. (11) of the main text follows. Note that our theory can as well describe the more general case in which presynaptic spike trains Sj are non-Poisson, but have stationary covariance function Cji (s, u) = Cji (s − u) and stationary mean hSj (u)i = νpre. When we insert (S4) instead of (S5) into (S3) we obtain the more general expression for the expected correlation term at the synaptic contact hCjki = hrjk λi = (1 − pf)Xi with the eective spike train covariance [pfwik + (1 − pf)wi ] ¯Cji + νpre(1 − pf)hRposti , (S6) ¯ Cji = τ−2 ∞ 0 ∞ 0 e−(u+s)/τCji (u − s + d) du ds . Eq. (S6) shows that the plasticity model is generally sensitive to spike-timing correlations of neurons all across the network. S1.2 Mathematical analysis of system dynamics. As our simulations show, the learning rule maintains a constant postsynaptic ring rate νpost by tight regulation of the total input weight. Thus the actual degrees of freedom of the multi-connection multi- contact system reside in the conguration of which of the connections wj are strong or weak, and through which number of contacts wjk this strength is achieved. To understand these dynamics, we consider the is expected change of the weight of a single contact dhwjki/dt assuming the total input w = PN constant and given by Eq. (10) in the main text. This analysis is displayed in Fig. S1. With the parameters of the learning rule (Eq. (7) of the main text) our mathematical analysis shows that the steady state has the following properties: (a) There is a stable xed point of Eq. (9) (marked by the circle in Fig. S1), such that if only 10% of the synaptic connections are active (consistent with [1, 2]), with total weight wj = w∗ each, the postsynaptic ring rate of νpost = νpre = 5/s is achieved. From Eq. (10) in the main text follows that this xed point weight has to be w∗ = (νpost−λ0)/[(1−pf)νpre·10%·N] = 0.016. The creation weight wc (cross in Fig. S1) was set arbitrarily to 15% of the xed point weight of a contact, wc = 0.15 w∗/5 (our results are insensitive against the exact value). (b): Contacts only have a stable xed point of wjk in connections that contain at least three active contacts (dash-dotted lines in Fig. S1A). In connections with less than three active contacts (dashed lines in Fig. S1A), dhwjki/dt is generally smaller than 0, so that these contacts are removed eventually. If (b) is fullled, contacts in connections with about 5 contacts, as observed experimentally in [1], are stable. j=1 wj To calibrate our plasticity model (Eq. (7) of the main text) we used a numerical optimization procedure to nd parameters acorr and α that fulll these constraints. As the conguration of xed points of Eq. 9 of the main text is insensitive to a common rescaling of these parameters, we manually adjusted the scale of the parameters to achieve a good agreement of the temporal dynamics of the contacts (cf. Fig. 2A,D,E in the main text) with the experimental references. , apost , acorr 2 4 4 2 Figure S1: Expected dynamics of synaptic contacts. A: Positive or negative change (red or blue color, Eq. 9) of a synaptic contact as a function of its weight wjk (horizontal axis) and the total weight of the connection from presynaptic neuron j (wj , vertical axis), under the assumption that the sum of all weights of all presynaptic neurons is implicitly normalized via Eq. (10). Straight lines mark synapses where wj is a multiple of wjk , i.e. synapses consisting of 1 and 2 contacts (dashed), or 3, 5 and 10 contacts (dash-dotted). The circle marks the stable xed point (w∗/5, w∗) of the dynamics (for 5 actual contacts), the white cross marks the point of a newly created contact in a previously inactive connection (wc, wc). The black curve marks combinations of wjk and wj that have zero expected change (wjk nullcline). B-C: Expected trajectories after a perturbation. We assume that in a synaptic connection j with nj contacts, nj − 1 contacts have identical values while a single contact wjk is perturbed. B: In a synaptic connection with ve contacts (xed point at wj = w∗, circle), four contacts have the same weight w∗/5 while the contact wjk is perturbed by a small amount w∆. Its trajectory (red line) starts at (w∗/5 + w∆, w∗ + w∆) and evolves back towards the xed point (circle). C: Five contacts have existed for a long time and each take a value wjl = w∗/5 consistent with the xed point (circle), when spontaneously a new contact wjk is created at its creation value wc. The trajectory of the new contact (red line) starts at (wc, w∗ + wc) and evolves towards the new xed point for 6 contacts. The previously existing contacts also evolve towards the new xed point (green line). S1.3 Expected trajectories in response to perturbations. The expected dynamics (Eq. (9) in the main text) allow us to predict trajectories of (wjk , wj ). In Fig. S1B we show that, in expectation, if a contact weight wjk in an active connection at steady-state is perturbed, the perturbation decays. In Fig. S1C, we display what happens when a new contact is created at the steady state in an active connection with ve contacts. We nd that the new contact as well as the ve existing contacts move to a new xed point corresponding to six stable contacts. As we have seen above, a synaptic connetion consisting of three or more contacts is stable in expectation. In Fig. S2 we consider the creation of synaptic contacts in an inactive connection, in which no other active contacts exist. Both for a single newly created contact (Fig. S2A) and for two simultaneously created new contacts (Fig. S2B), the new contacts are expected to approach zero. A phase plane analysis of the dynamics of two contacts (Fig. S2C) further conrms that, in expectation, all connections of only two contacts are eventually removed because of the absence of stable xed points. However, because of a region on the phase plane where the dynamics are slow, a connection with two contacts has a lifetime that is suciently long so that occasionally a third contact may be added. Still, due to the low creation rate we expect this event to be rare. These theoretical considerations suggest that the system is indeed calibrated to have a steady state that is qualitatively consistent with the experimental contact number distributions, in the sense that stable connections have at least three synaptic contacts. S1.4 Analytical integration of synaptic dynamics. As we show here, the dynamics of the plasticity model can be integrated analytically in the absence of spikes. This allows us to iterate the dynamics from spike to spike precisely and eciently, and to implement it as an event-driven algorithm in NEST [3]. If for all t ∈ (t0, t1) there are no transmitted pre- nor postsynaptic spikes, Sjk (t) = Spost(t) = 0, then the traces dened in Eqs. (3), (4) and (6) in the main text decay exponentially as rjk (t) = rjk (t0) exp(−(t − t0)/τ ), rpost(t) = rpost(t0) exp(−(t − t0)/τ ) and Rpost(t) = Rpost(t0) exp(−(t − t0)/τslow), respectively. 3 ai ebi (t−t0) , (S8) 7 X i=1 1 c wjk (t) = Figure S2: Transient synaptic contacts in inactive connections. A: New contact in a connection without other actual contacts. The new contact approaches wjk = 0 and is removed (red line). B: Simultaneous creation of two new contacts in a connection without other contacts. Both of the contacts identically approach wjk = 0 and are removed (red line). C: Phase plane analysis of the two contacts weights wjk and wjl for a connection made of 2 contacts (k 6= l). The nullclines of wjk (green) and wjl (red) do not intersect, therefore the system does not have a xed point, but there is a region where the nullclines are close to each other where the ow is rather slow. Colored arrows denotes the gradient d/dt(wjk , wjl ) on the nullclines (arrow length corresponds to gradient magnitude in arbitrary units). Black streamlines indicate direction of ow of the dynamics, thickness of arrows denotes speed. Visibly, all trajectories converge to (0, 0), yet the ghosts of an unstable xed point in the region where nullclines are close slows the dynamics suciently to allow for the occasional creation of additional contacts. Inserting these functions into Eq. (5) of the main text and solving the dierential equation yields Cjk (t) = Cjk (t0)e−(t−t0)/τslow + rjk (t0)rpost(t0) e−2(t−t0)/τ − e−(t−t0)/τslow 1 − 2τslow/τ . (S7) Similarly also wjk (t) can be solved for analytically. We insert the solution for Rpost(t) and Eq. (S7) and solve Eq. (7) of the main text. The resulting analytical expression for wjk (t) is a generalized Dirichlet polynomial given by with the parameters a1 = 2acorr 4 rjk (t0)rpost(t0)τ 2(−4 + ατ )(−2 + ατ )τslow(−(Cjk,post(t0)τ ) + rjk (t0)rpost(t0)τ + 2Cjk,post(t0)τslow)(−4 + ατslow)(−2 + ατslow)(−1 + ατslow) , a2 = acorr 2 (−4 + ατ )(−2 + ατ )(−(rjk (t0)rpost(t0)τ ) + Cjk,post(t0)(τ − 2τslow))(τ − 2τslow)τslow(−4 + ατslow) · (−2 + ατslow)(−2τslow + τ (−1 + ατslow)) , a3 = −(acorr 4 (−4 + ατ )(−2 + ατ )τslow(−(Cjk,post(t0)τ ) + rjk (t0)rpost(t0)τ + 2Cjk,post(t0)τslow)(−4 + ατslow)(−1 + ατslow)(−2τslow + τ (−1 + ατslow))) , a4 = −(apost 4 R 4 post(t0)(−4 + ατ )(−2 + ατ )(τ − 2τslow)2τslow · (−2 + ατslow)(−1 + ατslow)(−2τslow + τ (−1 + ατslow))) , a5 = −(acorr 4 r 2 jk (t0)r 2 post(t0)τ 3(−2 + ατ )(−4 + ατslow) · (−2 + ατslow)(−1 + ατslow)(−2τslow + τ (−1 + ατslow))) , 4 a6 = acorr 2 rjk (t0)rpost(t0)τ 2(−4 + ατ )(τ − 2τslow)(−4 + ατslow) · (−2 + ατslow)(−1 + ατslow)(−2τslow + τ (−1 + ατslow)) , a7 = (τ − 2τslow)2(wjk (t0)(−4 + ατ )(−2 + ατ )(−4 + ατslow)(−2 + ατslow) · 2 (−1 + ατslow)(−2τslow + τ (−1 + ατslow)) + acorr (−2 + ατslow)(rjk (t0)rpost(t0)τ + Cjk,post(t0)(2 − ατ )τslow) · (−2τslow + τ (−1 + ατslow)) + (−2 + ατ )(−1 + ατslow)(apost (−4 + ατ )τslow(−2 + ατslow)(−τ − 2τslow + ατ τslow) + acorr (2r 2 2rjk (t0)rpost(t0))τ (−4 + ατ )τslow + C 2 post(t0)τ 2 − Cjk,post(t0)(Cjk,post(t0) + jk (t0)r 2 4 jk,post(t0)(−4 + ατ )(−2 + ατ )τ 2 slow))) , (−4 + ατ )(−4 + ατslow) · 4 R 4 post(t0) · (−4 + ατslow) · b1 = −(1/τslow + 2/τ ), −2/τ , and b7 = −α, b2 = −1/τslow, b3 = −2/τslow, b4 = −4/τslow, b5 = −4/τ , b6 = c = (−4 + ατ )(−2 + ατ )(τ − 2τslow)2(−4 + ατslow) · (−2 + ατslow)(−1 + ατslow)(−2τslow + τ (−1 + ατslow)) . We can thus integrate the dynamics of the plasticity model exactly between any two spikes by evaluating Eq. (S8). At any time t at which there is a spike (Sjk (t) 6= 0 or Spost(t) 6= 0), we merely have to increment the traces: rjk (t) ← rjk (t) + 1/τ , rpost(t) ← rpost(t) + 1/τ , or Rpost(t) ← Rpost(t) + 1/τslow, respectively. Then the solutions of the dierential equations for the subsequent inter-spike-interval are as given above, with t0 ← t. When iterating the time-evolution of wjk from spike to spike by Eq. (S8), however, we need to make sure that wjk (t) does not cross the zero line at any intermediate t ∈ (t0, t1) (zero-crossings are important because they cause pruning of the contact). Therefore we apply Theorem 4.7 of [4], which provides a criterion to rule out zero-crossings of wjk (t) for any t > t0. If zero crossings cannot be ruled out by this criterion, we evaluate Eq. (S8) for all t in the interval (t0, t1), on the simulation time grid ∆t, to guarantee that no zero-crossings of wjk were missed. References [1] Markram H, Lübke J, Frotscher M, Roth A, Sakmann B (1997) Physiology and anatomy of synaptic connections between thick tufted pyramidal neurones in the developing rat neocortex. J Physiol 500 ( Pt 2): 409440. [2] Fares T, Stepanyants A (2009) Cooperative synapse formation in the neocortex. Proc Natl Acad Sci U S A 106: 1646316468. [3] Bos H, Morrison A, Peyser A, Hahne J, Helias M, et al. (2015). Nest 2.10.0. doi:10.5281/zenodo.44222. URL http://dx.doi.org/10.5281/zenodo.44222. [4] Jameson GJO (2006) Counting zeros of generalised polynomials: Descartes rule of signs and laguerres extensions. Math Gazette 90: 223-234. 5
1906.09290
1
1906
2019-06-21T18:46:09
Brain state stability during working memory is explained by network control theory, modulated by dopamine D1/D2 receptor function, and diminished in schizophrenia
[ "q-bio.NC", "physics.bio-ph" ]
Dynamical brain state transitions are critical for flexible working memory but the network mechanisms are incompletely understood. Here, we show that working memory entails brainwide switching between activity states. The stability of states relates to dopamine D1 receptor gene expression while state transitions are influenced by D2 receptor expression and pharmacological modulation. Schizophrenia patients show altered network control properties, including a more diverse energy landscape and decreased stability of working memory representations.
q-bio.NC
q-bio
Title Page Brain state stability during working memory is explained by network control theory, modulated by dopamine D1/D2 receptor function, and diminished in schizophrenia Urs Braun1,2* MD, Anais Harneit1 MSc, Giulio Pergola3 PhD, Tommaso Menara4 MSc, Axel Schaefer5 PhD, Richard F. Betzel6 PhD, Zhenxiang Zang1 MSc, Janina I. Schweiger1 MD, Kristina Schwarz1 MSc, Junfang Chen1 MSc, Giuseppe Blasi3 MD PhD, Alessandro Bertolino3 MD PhD, Daniel Durstewitz7 PhD, Fabio Pasqualetti4 PhD, Emanuel Schwarz1 PhD, Andreas Meyer-Lindenberg1 MD, Danielle S. Bassett2,8# PhD, Heike Tost1# MD 1 Department of Psychiatry and Psychotherapy, Central Institute of Mental Health, Medical Faculty Mannheim, University of Heidelberg, Mannheim, Germany 2 Department of Bioengineering, University of Pennsylvania, Philadelphia, PA, USA 3 Department of Basic Medical Science, Neuroscience, and Sense Organs, University of Bari Aldo Moro, 70124 Bari, Italy 4 Mechanical Engineering Department, University of California at Riverside, Riverside, CA, USA 5 Bender Institute of Neuroimaging, Justus Liebig University Giessen & Center for Mind, Brain and Behavior, University of Marburg and Justus Liebig University Giessen, Giessen, Germany 6 Department of Psychological and Brain Sciences, Indiana University, Bloomington, IN, USA 7 Department of Theoretical Neuroscience, Central Institute of Mental Health, Medical Faculty Mannheim/Heidelberg University, 68159 Mannheim, Germany 8 Department of Psychiatry, Department of Neurology, Department of Physics & Astronomy, and Department of Electrical & Systems Engineering, University of Pennsylvania, Philadelphia, PA, USA # these authors contributed equally Corresponding Author: Urs Braun M.D., Department of Psychiatry and Psychotherapy, Central Institute of Mental Health, Medical Faculty Mannheim, University of Heidelberg, J5, 68159 Mannheim, Germany; Tel.: +49 621 1703 6519; e-mail: [email protected] Key Words: network control theory, executive function, working memory, schizophrenia, dopamine Short title: Control over brain states during working memory impacted by dopamine function. Word Count (unreferenced Abstract): 72/70 Word Count (Article Body): 1488/1500 Number of Figures: 2/3 1 Number of References: 20/20 This paper contains Supplementary Materials. 2 Dynamical brain state transitions are critical for flexible working memory but the network mechanisms are incompletely understood. Here, we show that working memory entails brain- wide switching between activity states. The stability of states relates to dopamine D1 receptor gene expression while state transitions are influenced by D2 receptor expression and pharmacological modulation. Schizophrenia patients show altered network control properties, including a more diverse energy landscape and decreased stability of working memory representations. Working memory is an essential part of executive cognition depending on prefrontal neurons functionally modulated through dopamine D1 and D2 receptor activation (1-3). The dual-state theory of prefrontal dopamine function links the differential activation of dopamine receptors to two discrete dynamical regimes: a D1-dominated state with a high energy barrier favoring robust maintenance of cognitive representations and a D2-dominated state with a flattened energy landscape enabling flexible switching between states (4). Recent accounts extend the idea of dopamine's impact on working memory from a local prefrontal to a brain-wide network perspective (5, 6), but the underlying neural dynamics and brain-wide interactions have remained unclear. Network control theory (NCT) can be used to model brain network dynamics as a function of interconnecting white matter tracts and regional control energy (7). Based on the connectome, NCT can be used to examine the landscape of brain activity states: that is, which states within a dynamic scheme would the system have difficulty accessing, and more importantly, which regions need to be influenced (and to what extent) to make those states accessible (8). Specifically, to quantify accessibility, we approximate brain dynamics locally by a simple linear dynamical system, 𝑥(𝑡)=𝑨𝑥(𝑡)+𝑩𝑢(𝑡), where x (t) is the brain state inferred from functional magnetic resonance imaging (fMRI), A is a structural connectome inferred from DTI data, u is the control input, and B is a matrix describing which regions enact 3 control. To investigate, based on this conception, how the brain transitions between different cognitive states, we defined states as individual brain activity patterns related to a working memory condition (2-back) and to an attention control condition requiring motor response (0- back) in a sample of 178 healthy individuals undergoing fMRI (Fig. S1; Online Methods). Further, we obtained individual structural connectomes from white matter by DTI fiber tracking, and computed the optimal control energy necessary to drive the dynamical system from the 0-back activity pattern to the 2-back pattern, or vice versa (Fig. S2). We defined the stability of both brain states as the inverse energy necessary to revisit that state, where the energy, loosely, is defined as the average size of the control signals u(t) needed to instantiate a specific trajectory in the dynamical system as defined above (see Eq. 3 & 5 in Online Methods). As expected, the cognitively more demanding 2-back state was less stable (i.e., required higher energy for maintenance) than the control state (Fig. 1a; repeated measures ANOVA: main effect of 0- vs. 2-back stability: F(1,173) = 66.80, p < 0.001, see Online Methods for details on all analyses). Further, the stability of the 2-back state was significantly associated with working memory accuracy (Fig. 1b; b = 0.274, p = 0.006), suggesting that more stable 2-back network representations support higher working memory performance. We next investigated how the brain flexibly changes its activity pattern between states. Transitioning into the cognitively more demanding 2-back state required more control energy than the opposite transition (Fig. 1c; repeated measures ANOVA: F(1,174) = 27.98, p = 0.001). Other analyses suggested that prefrontal and parietal cortices steer both types of transitions, while default mode areas are preferentially important for the switch to the more cognitively demanding state (Fig. 1d; Online Methods). These results are in line with the assumed role of frontal-parietal circuits in steering brain dynamics (9) and shifting brain connectivity patterns (10); they also emphasize the importance of the coordinated behavior of brain systems commonly displaying deactivations during demanding cognitive tasks (11). 4 Following from the dual-state theory of network function, the stability of task-related brain states should be related to prefrontal D1 receptor status. To estimate individual prefrontal D1 receptor expression, we utilized methods relating prefrontal cortex D1 and D2 receptor expression to genetic variation in their co-expression partner (Online Methods), thereby enabling us to predict individual dopamine receptor expression levels from genotype data across the whole genome (12, 13). We found that D1 (but not the D2) expression-related gene score predicted stability of both states (Fig. 2a; 0-back: b = 0.184, p = 0.034; 2-back: b = 0.242, p = 0.007, Online Methods), in line with the assumed role of D1-related signaling in maintaining stable activity patterns during task performance (4, 14). Independent of stability, switching between different activity representations should relate to dopamine D2 receptor function. Indeed, when controlling for stability as a nuisance covariate in the regression model, the control energy of both state transitions could be predicted by the D2 (but not the D1) receptor expression gene score (Fig. 2b; 0- to 2-back: b = -0.076, p = 0.037; and trending for 2- to 0-back: b = -0.134, p = 0.068, Online Methods). This finding is particularly interesting, as it suggests that the function of D1 and D2 receptors are differentially, but cooperatively, involved in steering brain dynamics between different activity patterns, in line with previous research on D1 and D2 functioning in prefrontal circuits (4, 15). Our results thus far support the notion that the brain is a dynamical system in which the stability of a state is substantially defined by cognitive effort and modulated by D1 receptor expression, while transitions between states depend primarily on D2 receptor expression. If true, such a system should be sensitive to dopaminergic manipulation, and interference with D2-related signaling should reduce the brain's ability to control its optimal trajectories, i.e. increase the control energy needed when switching between states. To test these hypotheses, we investigated an independent sample of healthy controls (n=16, Table S2) receiving 400 mg Amisulpride, a selective D2 receptor antagonist, in a randomized, placebo- controlled, double-blind pharmacological fMRI study. As expected, we observed that greater control energy was needed for transitions under D2 receptor blockade (Fig. 2c; repeated 5 measures ANOVA with drug and transition as within-subject factors; main effect of drug: F(1,10) = 7.27, p = 0.022; drug-by-condition interaction: F(1,10) = 0.42, p = 0.665). We observed no effect on the stability of states; that is, the inverse control energy required to stabilize a current state (main effect of drug: F(1,8) = 0.715, p = 0.422, Table S3). Dopamine dysfunction, working memory deficits, and alterations in brain network organization are hallmarks of schizophrenia (16-19). We therefore tested for differences in the state stability and in the ability to control state transitions between schizophrenia patients and a healthy control sample balanced for age, sex, performance, head motion, and premorbid IQ (see Table S1). Stability in schizophrenia patients was reduced for the cognitively demanding working memory state (F(1,98) = 6.43, p = 0.013), but not for the control condition (F(1,98) = 0.052, p = 0.840, Table S3). Control energy needed for the 0- to 2-back transition was significantly higher in schizophrenia (Fig. 2d; F(1,98) = 5.238, p = 0.024), while the opposite transition showed no significant group difference (ANOVA: F(1,98) = 0.620, p = 0.433, Table S3), in line with clinical observations that D2 blockade does not ameliorate cognitive symptoms in schizophrenia (20). These results suggest that the brain energy landscape is more diverse in schizophrenia, making the system more difficult to steer appropriately. To further strengthen this notion, we estimated the variability in suboptimal (higher energy) trajectories connecting different of cognitive states (Online Methods). We expected that in a diversified energy landscape, the variation of trajectories around the minimum-energy trajectory should be larger, implying that small perturbations may have a more substantial impact. In line with our hypothesis, we found that the variability in such perturbed trajectories was indeed increased in schizophrenia (rm-ANOVA: main effect of group: F(1,98) = 4.789, p = 0.031, Online Methods). Several aspects of our work require special consideration. Firstly, to relate brain dynamics to cognitive function, we focus on discrete brain states where each state is summarized by a single brain activation patterns rather than linear combination of multiple brain activity patterns. Secondly, although we could demonstrate a link between brain dynamics, 6 measured by means of control energy, and predicted prefrontal dopamine receptor expression, the link is indirect and requires confirmation by direct measurements. Thirdly, we cannot exclude the possibility that disorder severity, duration, symptoms or medication may have influenced network dynamics in schizophrenia patients, although our supplemental analyses do not support this conclusion (Online Methods). Finally, while the sample sizes of our pharmacological and patient study are rather small, we were able to show comparable effects of dopaminergic manipulation on control properties using a second (Online Methods), further supporting the validity of the underlying rationale. In summary, our data demonstrate the utility of network control theory for the non-invasive investigation of the mechanistic underpinnings of (altered) brain states and their transitions during cognition. Our data suggest that engagement of working memory involves brain-wide switching between activity states and that the steering of these network dynamics is differentially, but cooperatively, influenced by dopamine D1 and D2 receptor function. Moreover, we show that schizophrenia patients show reduced controllability and stability of working memory network dynamics, consistent with the idea of an altered functional architecture and energy landscape of cognitive brain networks. 7 Acknowledgements The authors thank all individuals who have supported our work by participating in our studies. There was no involvement by the funding bodies at any stage of the study. We thank Oliver Grimm, Leila Haddad, Michael Schneider, Natalie Hess, Sarah Plier and Petya Vicheva for valuable research assistance. The authors thank Jason Kim and Lorenzo Caciagli for valuable feedback on the manuscript. U.B. acknowledges grant support by the German Research Foundation (DFG, grant BR 5951/1-1). H.T. acknowledges grant support by the German Research Foundation (DFG, Collaborative Research Center SFB 1158 subproject B04, Collaborative Research Center TRR 265 subproject A04, GRK 2350 project B2, grant TO 539/3-1) and German Federal Ministry of Education and Research (BMBF, grants 01EF1803A project WP3, 01GQ1102). AML acknowledges grant support by the German Research Foundation (DFG, Collaborative Research Center SFB 1158 subproject B09, Collaborative Research Center TRR 265 subproject S02, grant ME 1591/4-1) and German Federal Ministry of Education and Research (BMBF, grants 01EF1803A, 01ZX1314G, 01GQ1003B), European Union's Seventh Framework Programme (FP7, grants 602450, 602805, 115300 and HEALTH-F2-2010- 241909, Innovative Medicines Initiative Joint Undertaking (IMI, grant 115008) and Ministry of Science, Research and the Arts of the State of Baden-Wuerttemberg, Germany (MWK, grant 42-04HV.MED(16)/16/1). DSB and RBF would like to acknowledge support from the John D. and Catherine T. MacArthur Foundation, the Alfred P. Sloan Foundation, the Army Research Laboratory and the Army Research Office through contract numbers W911NF-10-2-0022 and W911NF-14-1-0679, the National Institute of Health (2-R01-DC-009209-11, 1R01HD086888- 01, R01-MH107235, R01-MH107703, and R21-M MH-106799), the Office of Naval Research, and the National Science Foundation (BCS-1441502, CAREER PHY-1554488, and BCS- 1631550). E.S. gratefully acknowledges grant support by the Deutsche Forschungsgemeinschaft, DFG (SCHW 1768/1-1). X.L.Z. is a Ph.D. scholarship awardee of the Chinese Scholarship Council. DD acknowledges grant support by the German Research 8 Foundation (DFG, Du 354/10-1). G.P. has received funding from the European Union's Horizon 2020 research and innovation program under the Marie Skłodowska-Curie No. 798181: "IdentiFication of brain deveLopmental gene co-expression netwOrks to Understand RIsk for SchizopHrenia" (FLOURISH). The content of this paper is solely the responsibility of the authors and does not necessarily represent the official views of any of the funding agencies 9 Financial disclosures A.M.-L. has received consultant fees from Blueprint Partnership, Boehringer Ingelheim, Daimler und Benz Stiftung, Elsevier, F. Hoffmann-La Roche, ICARE Schizophrenia, K. G. Jebsen Foundation, L.E.K Consulting, Lundbeck International Foundation (LINF), R. Adamczak, Roche Pharma, Science Foundation, Synapsis Foundation -- Alzheimer Research Switzerland, System Analytics, and has received lectures including travel fees from Boehringer Ingelheim, Fama Public Relations, Institut d'investigacions Biomèdiques August Pi i Sunyer (IDIBAPS), Janssen-Cilag, Klinikum Christophsbad, Göppingen, Lilly Deutschland, Luzerner Psychiatrie, LVR Klinikum Düsseldorf, LWL PsychiatrieVerbund Westfalen-Lippe, Otsuka Pharmaceuticals, Reunions i Ciencia S. L., Spanish Society of Psychiatry, Südwestrundfunk Fernsehen, Stern TV, and Vitos Klinikum Kurhessen. A.B. has received consultant fees from Biogen and speaker fees from Lundbeck, Otsuka, Recordati, and Angelini. The remaining authors reported no biomedical financial interests of potential conflicts of interest. 10 References Kim JZ, Soffer JM, Kahn AE, Vettel JM, Pasqualetti F, Bassett DS (2018): Role of Graph Cole MW, Reynolds JR, Power JD, Repovs G, Anticevic A, Braver TS (2013): Multi-task Durstewitz D, Seamans JK (2008): The dual-state theory of prefrontal cortex dopamine Arnsten AF (2011): Catecholamine influences on dorsolateral prefrontal cortical networks. Roffman JL, Tanner AS, Eryilmaz H, Rodriguez-Thompson A, Silverstein NJ, Ho NF, et al. Betzel RF, Gu S, Medaglia JD, Pasqualetti F, Bassett DS (2016): Optimally controlling the Ferenczi EA, Zalocusky KA, Liston C, Grosenick L, Warden MR, Amatya D, et al. (2016): Meyer-Lindenberg A, Kohn PD, Kolachana B, Kippenhan S, McInerney-Leo A, Nussbaum R, et Goldman-Rakic PS (1995): Cellular basis of working memory. Neuron. 14:477-485. Ott T, Jacob SN, Nieder A (2014): Dopamine receptors differentially enhance rule coding in 1. 2. primate prefrontal cortex neurons. Neuron. 84:1317-1328. 3. al. (2005): Midbrain dopamine and prefrontal function in humans: interaction and modulation by COMT genotype. Nat Neurosci. 8:594-596. 4. function with relevance to catechol-o-methyltransferase genotypes and schizophrenia. Biol Psychiatry. 64:739-749. 5. Biol Psychiatry. 69:e89-99. 6. (2016): Dopamine D1 signaling organizes network dynamics underlying working memory. Sci Adv. 2:e1501672. 7. Architecture in Controlling Dynamical Networks with Applications to Neural Systems. Nat Phys. 14:91-98. 8. human connectome: the role of network topology. Sci Rep. 6:30770. 9. Prefrontal cortical regulation of brainwide circuit dynamics and reward-related behavior. Science. 351:aac9698. 10. connectivity reveals flexible hubs for adaptive task control. Nat Neurosci. 16:1348-1355. 11. Greicius MD, Krasnow B, Reiss AL, Menon V (2003): Functional connectivity in the resting brain: a network analysis of the default mode hypothesis. Proceedings of the National Academy of Sciences of the United States of America. 100:253-258. 12. Fazio L, Pergola G, Papalino M, Di Carlo P, Monda A, Gelao B, et al. (2018): Transcriptomic context of DRD1 is associated with prefrontal activity and behavior during working memory. Proc Natl Acad Sci U S A. 115:5582-5587. 13. expression network and a related polygenic index predict imaging, behavioral and clinical phenotypes linked to schizophrenia. Transl Psychiatry. 7:e1006. 14. Dopaminergic modulation of distracter-resistance and prefrontal delay period signal. Psychopharmacology. 232:1061-1070. 15. differential D1 versus D2 dopamine receptor regulation of inhibition in prefrontal cortex. J Neurosci. 24:10652-10659. 16. common pathway. Schizophrenia bulletin. 35:549-562. 17. CNTRICS and schizophrenia. Biol Psychiatry. 64:11-17. 18. Tost H, Alam T, Meyer-Lindenberg A (2010): Dopamine and psychosis: theory, pathomechanisms and intermediate phenotypes. Neurosci Biobehav Rev. 34:689-700. 19. Braun U, Schafer A, Bassett DS, Rausch F, Schweiger JI, Bilek E, et al. (2016): Dynamic brain network reconfiguration as a potential schizophrenia genetic risk mechanism modulated by NMDA receptor function. Proc Natl Acad Sci U S A. 113:12568-12573. 20. Millan MJ, Fone K, Steckler T, Horan WP (2014): Negative symptoms of schizophrenia: clinical characteristics, pathophysiological substrates, experimental models and prospects for improved Bloemendaal M, van Schouwenburg MR, Miyakawa A, Aarts E, D'Esposito M, Cools R (2015): Pergola G, Di Carlo P, D'Ambrosio E, Gelao B, Fazio L, Papalino M, et al. (2017): DRD2 co- Trantham-Davidson H, Neely LC, Lavin A, Seamans JK (2004): Mechanisms underlying Howes OD, Kapur S (2009): The dopamine hypothesis of schizophrenia: version III--the final Barch DM, Smith E (2008): The cognitive neuroscience of working memory: relevance to 11 treatment. European neuropsychopharmacology : the journal of the European College of Neuropsychopharmacology. 24:645-692. 12 Figures Figure 1: Controllability and stability of brain dynamics during working memory A) The stability of the 2-back state reflecting working memory activity is lower than that of the 0-back state reflecting motor and basic attention control activity (F(1,173) = 66.80, p < 0.001). Red lines indicate mean values and boxes indicate one standard deviation of the mean. B) Associations of 2-back stability with working memory performance (accuracy: b = 0.274, p = 0.006; covarying for age, sex, and mean activity). C) Steering brain dynamics from the control condition to the working memory condition requires more control energy than vice versa (F(1,174) = 27.98, p < 0.001). D) Unique and common sets of brain regions contribute most to the transition from 0-back to 2-back and the transition from 2-back to 0-back transitions, respectively. For illustrative purposes, we projected the computed control impact 13 of each brain region (Online Methods) for the respective transitions on a 3D structural template, displaying the 20% highest for each transition. Figure 2: Dopamine receptor expression and pharmacological modulation impact whole brain dynamics A) Genetic scores predicting DRD1 expression in prefrontal regions positively predict stability of both brain states (0-back: b = 0.184, p = 0.034; 2-back: b = 0.242 p = 0.007; age, sex, mean brain state activity, first 5 genetic PCA components as covariates of non-interest). B) Genetic scores predicting DRD2 expression in prefrontal regions negatively predict control energy for both brain state transitions (0-back to 2-back: b = -0.076, p = 0.037; and trend wise for 2-back to 0-back: b = -0.134, p = 0.068; age, sex, mean brain activity difference, first 14 5 genetic PCA components, stability of 0-back and 2-back as covariates of non-interest). C) Amisulpride increases control energy for transitions in comparison to placebo (main effect of drug: F(1,10) = 7.27, p = 0.022; interaction drug by condition: F(1,10) = 0.42, p = 0.665, activity difference, drug order, and sex as covariates of non-interest). Black lines indicate mean values and boxes indicate one standard deviation of the mean. D) Schizophrenia patients need more control energy when transitioning into the working memory condition than matched healthy controls (F(1,98) = 5.238, p = 0.024, age, sex, tSNR and mean activity as covariates of non-interest), but not vice versa. 15 ONLINE METHODS Urs Braun et al. "Brain state stability during working memory is explained by network control theory, modulated by dopamine D1/D2 receptor function, and diminished in schizophrenia" 6.1. 6.2. 6.3. 7.1. 7.2. 2.1. 2.2. 1. Participants and study design ....................................................................................................... 2 2. Data acquisition .............................................................................................................................. 3 fMRI ...................................................................................................................................................... 3 Diffusion Tensor Imaging (DTI) ............................................................................................................... 4 3. Atlas construction .......................................................................................................................... 4 4. Connectome construction .............................................................................................................. 5 5. Brain state definition ...................................................................................................................... 5 6. Network control theory ................................................................................................................... 6 Optimal control theory framework ......................................................................................................... 6 Control energy, stability, and impact ..................................................................................................... 7 Suboptimal trajectories ......................................................................................................................... 8 7. Gene based polygenic co-expression indices .............................................................................. 9 Genotyping, imputation and quality control ........................................................................................... 9 Polygenic co-expression index calculation ............................................................................................ 9 8. Statistical inference ...................................................................................................................... 11 9. Control analyses ........................................................................................................................... 12 Null models of structural brain networks .............................................................................................. 12 Null models of spatial activity patterns ................................................................................................ 12 Robustness to choice of parcellation scheme ..................................................................................... 12 Robustness to choice of edge definition .............................................................................................. 13 Impact of medication and duration of illness on control properties ...................................................... 13 Pharmacological validation using Risperidone ..................................................................................... 13 Null results for gene score and imaging associations .......................................................................... 14 Suboptimal trajectories ....................................................................................................................... 14 Supplementary figures ............................................................................................................. 15 10.1. Figure S1: N-back task design ............................................................................................................ 15 10.2. Figure S2: Network control theory concept and methods .................................................................... 16 Supplementary tables ............................................................................................................... 17 11.1. Table S1: Characteristics of the healthy control and patient samples .................................................. 17 11.2. Table S2: Sample characteristics of the pharmacological study .......................................................... 18 11.3. Table S3: Statistical details for the main findings and replication analyses .......................................... 19 11.4. Table S4: List of brain regions included in the extended Glasser parcellation ...................................... 21 Supplementary references ....................................................................................................... 24 9.1. 9.2. 9.3. 9.4. 9.5. 9.6. 9.7. 9.8. 10. 11. 12. 1 1. Participants and study design All participants provided written informed consent for protocols approved by the Institutional Review Board of the medical faculty in Mannheim. For the first study including healthy controls and patients with schizophrenia, a total of 202 subjects (178 healthy controls, 24 schizophrenia patients) were included (see Table S1). General exclusion criteria in controls included the presence of a lifetime history of psychiatric, neurological, or significant general medical illness, pregnancy, a history of head trauma, and current alcohol or drug abuse. The patients were recruited from the Department of Psychiatry and Psychotherapy at the Central Institute of Mental Health in Mannheim and via local advertisements. A trained psychiatrist or psychologist verified the diagnosis of schizophrenia based on ICD-10 criteria. Neuropsychological characterization of healthy controls included the trail-making-test B (TMT-B) (1) and the German multiple-choice vocabulary intelligence test (MWT-B) (2) as a measure of premorbid IQ. Clinical characterization included the assessment of current symptom severity using the Positive and Negative Symptom Scale (PANSS) (3), Beck's Depression Inventory (4), measures of global functioning in daily life (global assessment of function (GAF)) and current antipsychotic medication dosage (converted into chlorpromazine dose equivalents (CPZE)). For the second, pharmacological intervention study, 17 healthy individuals completed a subject- and observer-blind, placebo-controlled, randomized three-period cross-over study (see Table S2). Exclusion criteria included a regular consumption of drugs or history of drug or alcohol abuse; systolic blood pressure (SBP) greater than 140 or less than 90 mm Hg, and diastolic blood pressure (DBP) greater than 90 or less than 50 mm Hg; notable resting bradycardia (heart rate (HR) <40 bpm) or tachycardia (HR >90 bpm); use of any medication or herbal remedies taken within 14 days prior to randomization into the study or 5 times the elimination half-life of the medication, clinically significant abnormalities in laboratory test results (including hepatic and renal panels, complete blood count, chemistry panel and urinalysis): a history or presence of clinically significant ECG abnormalities (e.g. PQ/PR interval >210 ms, QTcF >450 ms) or cardiovascular disease (e.g. cardiac insufficiency, coronary artery disease, cardiomyopathy, hypokalemia, congestive heart failure, family history of congenital long QT syndrome, family history of sudden death); any personal or familial 2 history of seizures, epilepsy or other convulsive condition, previous significant head trauma, or other factors predisposing to seizures; disorders of the central nervous system, cerebrovascular events, Parkinson's disease, migraine, depression, bipolar disorder, anxiety, any other psychiatric disorders or behavioral disturbances; regular smoking (>5 cigarettes, >3 pipe-fulls, >3 cigars per day); habitual caffeine consumption of more than 400 mg/d (approximately 4 cups of coffee or equivalent); a history or evidence of any clinically significant endocrinological, hepatic, renal, autoimmune, pulmonary, gastrointestinal, urogenital, oncological, hematological or any other disease; or a body mass index (BMI) of over 30 or below 22. Participants were invited for a fixed interval of 7 days with each scanning session taking place at approximately the same time of day. On each of three scanning visits, individuals either received a single oral dose of 400 mg Amisulpride, 3 mg Risperidone or Placebo. MRI scanning took place 2 hours after drug administration, with the N-back paradigm commencing approximately 10 min after the start of the scan. One subject was excluded from the analysis due to an excessive body-mass index (BMI > 30). 2. Data acquisition 2.1. fMRI For the first study, BOLD fMRI was performed on a 3T Siemens Trio (Erlangen, Germany) in Mannheim, Germany. Prior to the acquisition of functional images, a high-resolution T1-weighted 3D MRI sequence was conducted (MPRAGE, slice thickness=1.0 mm, FoV = 256 mm, TR = 1570ms, TE = 2.75 ms, TI = 800ms, α = 15°). Subsequently, functional data was acquired during performance of the N-back paradigm using an echo-planar imaging (EPI) sequence with the following scanning parameters: TR/ TE = 2000/30 ms, α =80°, 28 axial slices (slice-thickness = 4 mm + 1 mm gap), descending acquisition, FoV = 192 mm, acquisition matrix = 64 x 64 128 volumes. For the second study, BOLD fMRI was performed on a 3T Siemens Trio Scanner (Erlangen, Germany) using an echo-planar imaging (EPI) sequence with the following parameters: TR = 1790 ms, TE = 28 ms, 34 axial slices per volume, voxel size = 3 x 3 x 3 mm, 1 mm gap, 192 x 192 mm field of view, 76° flip angle, descending acquisition. 3 The visual N-back paradigm is a well-established and reliable working memory task consisting of a high memory load (2-back) and an attention control condition requiring motor response (0-back) (5-7). Specifically, a diamond-shaped stimulus containing a number from 1 to 4 was presented every 2 seconds (see Figure S1). In the 0-back condition, subjects were required to press the button on the response box corresponding to the number currently displayed on the presentation screen. In the 2-back condition, subjects were required to press the button on the response box corresponding to the number presented two stimuli before the number currently displayed on the presentation screen. Stimuli were presented in alternating blocks of either 0-back or 2-back conditions. In each condition block, 14 stimuli were presented. Each condition block was repeated 4 times. Task performance was measured by accuracy (defined as the percent of correct answers) and reaction time (defined as the time span between stimulus onset and button press) for each condition separately. 2.2. Diffusion Tensor Imaging (DTI) DTI data were acquired by using spin echo EPI sequences with the following parameters: TR 14000ms, TE 86ms, 2mm slice thickness, 60 non-collinear directions, b-value 1000 s/mm2, 1 b0 image, FOV 256 mm. 3. Atlas construction To combine structural and functional brain imaging data, we first constructed a brain atlas that equally well respects functional and anatomical features. We transformed a recently published multimodal atlas (8) into a volumetric format by projecting its FreeSurfer pial cortex coordinates into standard MNI space. A grey matter prior probability map (thresholded at 0.3) as provided by SPM was used to define relevant voxels. Voxels were labeled by choosing the closest label with maximum distance of 4 mm. Since the published multimodal atlas does not cover all subcortical regions of interest (e.g. amygdala, thalamus), we complemented it with subcortical structures from the Harvard-Oxford atlas as implemented in FSL (9). Combining the two atlases resulted in 374 regions that covered cortical and subcortical structures. A full list of regions included in the combined atlas can be found in Supplemental Table S3. 4 4. Connectome construction For the DTI data, the following preprocessing steps were performed with standard routines implemented in the software package FSL (9): i) correction of the diffusion images for head motion and eddy currents by affine registration to a reference (b0) image, ii) extraction of non-brain tissues (10), and iii) linear diffusion tensor fitting. After estimation of the diffusion tensor, we performed deterministic whole-brain fiber tracking as implemented in DSI Studio using a modified FACT algorithm (11). For each subject, 1,000,000 streamlines were initiated. Streamlines with a length of less than 10 mm were removed (12). For the construction of structural connectivity matrices, the brain was parcellated into 374 regions. To map these parcellations into subject space, we applied a nonlinear registration implemented in DSI Studio. We estimated the structural connectivity between any two regions of the atlas by using the mean fractional anisotropy values between respective brain regions. This procedure resulted in a weighted adjacency matrix A whose entries Aij reflect the strength of structural connectivity between two brain regions. 5. Brain state definition Because we were interested in investigating how the brain controls and transitions between global brain states underlying circumscribed cognitive processes (such as those supporting working memory, attention, and motor behavior), we defined brain states as stationary patterns of activity during execution of these processes. It is important to note that the temporal resolution of fMRI and the design of the N-back task limits the investigation of differential cognitive processes contributing to memory performance on fast time scales. Therefore, we cannot investigate the detailed temporal dynamics of working memory processes. However, our simplified design allows us to extract meaningful brain states elicited by a controlled cognitive process and therefore enables us to relate brain dynamics to cognitive function. Specifically, we defined individual brain states as spatial patterns of beta estimates associated with activity across brain regions of interest during both conditions of the N-back task (13). For that purpose, after standard preprocessing procedures in SPM12/8 (including realignment to the mean image, slice time correction, spatial normalization into standard stereotactic space, resampling to 3 mm isotropic voxels, and smoothing with an 8 mm full-width at half-maximum Gaussian Kernel) (5, 6, 14), we estimated standard 5 first-level general linear models for the N-back task, separately for each individual. Except for the SPM version, both studies followed the same preprocessing procedure. These GLM models included regressors for the 0-back and 2-back conditions of interest, as well as the 6 motion parameters as regressors of non- interest. To define the brain activity pattern associated with each condition of the task, we extracted GLM (beta) parameter estimates for the 0-back and 2-back conditions separately (13) and we averaged them across all voxels in each of the 374 regions without applying any threshold. This procedure yielded a 374 by 1 vector for each condition (x0-back, x2-back) per subject representing how strongly the BOLD response in back), respectively. These vectors (x0-back, x2-back) defined the final and target brain states for the following each brain region was associated with the working memory (2-back) or the motor control condition (0- network control analyses. 6. Network control theory 6.1. Optimal control theory framework To model the transition between 0-back and 2-back brain states, we used the framework of optimal control, following prior work (15-17) implemented in MATLAB. Based on individual brain states X=[x1,…xn] (in our case simplified to n = 2 states: 0-back and 2-back, see above) and a structural brain network A for each (1) ẋ(t) = A x(t) + B u(t) subject, we approximated the local brain dynamics by a linear continuous-time equation: to model the flow among task-related brain activity states. In the model, x(t) is the state of the system at time t, A is the wiring diagram of the underlying network, B denotes an input matrix defining the control nodes, and u(t) is the time-dependent control signal (17, 18). Note that while the initial state (x0) and the target state (xT) are empirically defined, any states of the system at other times are virtual intermediate steps in the trajectories of the state-space model. The problem of finding an optimal control energy u* that induces a trajectory from an initial state x0 to a target state xT reduces to the problem of finding an optimal solution to the minimization problem of the corresponding Hamiltonian: (2) min[H(p, x, u, t) = xTx + ñuTu + pT (Ax + Bu)] (15). 6 For simplicity, we set the input matrix B = IN´N, the identity matrix, allowing all brain regions to be independent controllers (16, 18). 6.2. Control energy, stability, and impact Control energy for each node ki, i = 1…m (m = total number of brain nodes), was defined as (3) 𝐸𝑘𝑖=∫ D𝑢∗𝑘𝑖(𝑡)D2𝑑𝑡 𝑇𝑡=0 , i.e. the squared integral over time of energy input that the node has to exhibit to facilitate the transitions from the initial state to the target state (15, 16, 18). While the neurobiological foundations of control properties in the brain are not yet well understood, control energy can be interpreted as the effort of a brain region needed to steer the activity pattern of itself and its connected brain regions into the desired final activation state, for example by tuning their internal firing or activity patterns by recurrent inhibitory connections. Accordingly, the total control energy for the entire brain was defined as the sum of all control energy across all nodes (4) 𝐸=∑ 𝐸KL MNOP , yielding one value for each transition per subject. To ensure a normal distribution of metric values for subsequent statistical testing, we applied a logarithmic transformation (base 10) to the control energy(16). From the control energy, we can also obtain the control stability and control impact. Stability was defined as (5) 𝑆= P RSTUV(WXYZX[), i.e. the inverse control energy needed to maintain a state, or in other words, the control energy needed to go, e.g., from 2-back to 2-back (18). To investigate the influence that a single brain region has on the entire system's dynamics during state trajectories, we computed the control impact of each node by iteratively removing one brain region from the network and re-computing the change in control energy (15). 7 6.3. Suboptimal trajectories To investigate the energy landscape surrounding the minimum energy (in this sense 'optimal') trajectories, we quantified the variability of suboptimal trajectories by adding subtle random perturbation to the minimum energy trajectories over 200 iterations. Note that we employed a discrete-time dynamical system rather than a continuous-time system for these analyses, as discrete-time systems are computationally more tractable. To discretize our linear continuous-time system, we employed the following transformations (6) 𝑨] ≜ 𝑒𝑨'a and (7) 𝑩]≜ c∫ 𝑒d('aef) 'ag 𝑑𝜏i𝑩, where 𝑨] and 𝑩] are the corresponding structural matrix and control input matrix in a discrete time system and Ts is the sampling time. As there is no prescriptive way to choose Ts, we estimated 𝑇m= PPg 𝑅𝑇, where RT is the rise time of its fastest mode, i.e. the time that the system requires to go from 10% to 90% of its fastest step response. In the resulting discrete-time dynamical system (8) 𝑥 (𝑡+1)= 𝑨]𝑥(𝑡)+𝐵s𝑢s(𝑡), at each discrete time step t, we applied a principal component analysis (PCA) to the cloud of suboptimal points to reduce dimensionality. Because the previous simulations show that the first PCA component explains more than 90% of variance, we continued by using the first component as a summary measure of suboptimal trajectories. To estimate the relative distance between suboptimal and optimal trajectories, we computed the percentage of variation of the maximum distance of the projected suboptimal points on the first principal component from the optimal one. This procedure resulted in a normalized measure giving the percentage of deviations for the 200 suboptimal trajectories from the optimal trajectory. Due to the heuristic nature of the algorithm applying a random perturbation, results can vary from run to run. Therefore, to increase replicability of our results, we repeated our analysis 10 times per subject (10*200 = 2,000 suboptimal trajectories per subject). Because DTI data was not acquired during the pharmacological intervention study, we used the average connectivity matrix across all healthy subjects from study 1 to model the transition between individual brain states in the pharmacological intervention study. 8 7. Gene based polygenic co-expression indices 7.1. Genotyping, imputation and quality control In this study, we used human GWAS data of 63 healthy subjects who were genotyped using HumanHap 610 and 660w Quad BeadChips. For all subjects, standard quality control (QC) and imputation were performed using the Gimpute pipeline (Chen, Lippold et al. 2018) and the following established QC steps were applied. Step 1: Determine the number of male subjects for each heterozygous SNP of the chromosome 23 and remove SNPs whose number is larger than 5% of the number of male samples. Step 2: Determine the number of heterozygous SNPs on X chromosome for each male sample, and remove samples that have the number of heterozygous SNPs larger than 10. Step 3: Remove SNPs with missing genotyping rate > 5% before sample removal. Step 4: Exclude samples with missingness >= 0.02. Step 5: Exclude samples with autosomal heterozygosity deviation Fhet >= 0.2. Step 6: Remove SNPs with the proportion of missing genotyping > 2% after sample removal. Step 7: Remove SNPs if the Hardy-Weinberg equilibrium exact test P-value was < 1 x 10-6. Step 8: Principal component analysis (PCA) was applied to detect population outliers. Imputation was carried out using IMPUTE2/SHAPEIT (19-21), with a European reference panel for each study sample in each 3 Mb segment of the genome. This imputation reference set is from the full 1000 Genome Project dataset (August 2012, 30,069,288 variants, release "v3.macGT1"). The length of the buffer region is set to be 500 kb on either side of each segment. All other parameters were set to default values implemented in IMPUTE2. After imputation, SNPs with high imputation quality (INFO >= 0.6) and successfully imputed in >= 20 samples were retained. From the final well-imputed dataset with 63 subjects, we extracted 8 SNPs for DRD2 and 13 SNPs for DRD1 (22-24). For the subsequent genetic imaging analyses, we only used subjects for whom both data modalities (DTI and fMRI data) were available. 7.2. Polygenic co-expression index calculation Previous publications have shown that gene sets defined using co-expression networks and selected for their association with the genes DRD1 and DRD2 provided replicable predictions of n-back-related brain activity and behavioral indices (23, 25-27). Weighted Gene Co-expression Network Analysis [WGCNA (28)] applied 9 on the Braincloud dataset (N=199) of post-mortem DLPFC gene expression (29) identified 67 non- overlapping sets of genes based on their expression pattern. The co-expression gene sets including DRD1 and DRD2 were summarized into Polygenic Co-expression Indices (PCIs) based on SNPs that predicted co- expression of these genes (called co-expression quantitative trait loci, or co-eQTLs). PCIs are a proxy for the assessment of the genetic component of gene transcription co-regulation and are computed as a weighted average of the effect of all genotypes of an individual among those selected in the data mining study as co-eQTLs. The effect of individual SNPs is computed as the difference between the gene co-expression distribution of minor allele carriers (heterozygotes and homozygotes) and that of major allele homozygotes, using common tools from signal detection theory (30). Genotype weights, therefore, represent the deviation in gene co-expression from a reference distribution and are not constrained by allele dose. For each genotype of each SNP we computed an index, called A', proportional to the expression of the gene of interest (DRD1 or DRD2) within its co-expression module. The A' index is less dependent than d' on the assumption of a normal distribution of gene expression in each genotypic population (25). Both PCI-based predictions were significantly replicated in an independent post-mortem dataset, while controlling for ethnicity. The translational effect of these two scores on brain activity during n- back has been assessed and replicated across multiple samples, which combined amount to approximately 600 participants (22-24). It is important to note that these dopamine-related genetic effects are large in magnitude compared to those estimated by polygenic risk score approaches that focus on epidemiological data, rather than on molecular processes. The DRD2-PCI we developed (23) yielded an effect size f = 0.30 in our n-back discovery sample (required sample size to obtain 80% power with α = 0.05 and covariates as in the current work: N = 71). Results were replicated in an independent fMRI dataset collected at a different institution with f = 0.20 (required sample size computed as above: N = 156). Our follow-up work on the DRD2-PCI (26) considered two datasets of 50 individuals each and yielded a minimum effect size f = 0.28 (in the replication sample; required sample size computed as above: N = 81). The DRD1-PCI was also tested in two independent samples (25), yielding a minimum effect size f = 0.37 (in the replication sample; required sample size computed as above: N = 46). Taken together, these published results show that the effects of these polygenic indices on n-back activity in the prefrontal cortex are relatively large, with sizes ranging between 0.20 and 0.37 and with required samples ranging from 46 to 156 individuals. Importantly, the 10 DRD2-PCI was also tested in a small sample of 29 patients with SCZ and yielded results consistent with the effects discovered in healthy controls (23). Although the required sample sizes were computed based on the top cluster, it should be borne in mind that the technique we used in this work employs the entire brain, and therefore (i) is not subject to correction for multiple comparisons, as reflected in the uncorrected alpha used for the power calculations and (ii) benefits from the greatest possible amount of information about brain states. 8. Statistical inference Statistical inference was performed using the Statistical Package for the Social Sciences 24 (IBM SPSS Statistics for Windows, Version 24). All statistical comparisons were performed while controlling for age and sex and were tested two-sided. Because we were interested in control properties of brain state transitions independent of differences in the amount of mere activation, we controlled for the respective parameters reflecting the individual differences in activations in all analyses involving control properties. In particular, for all analyses involving stability measures, we additionally controlled for the average brain energy defined as the GLM parameter estimates over all regions in the 0-back and 2-back conditions. As the control energy of a transition depends highly on the absolute difference in energy between its initial state and its final state (i.e. in our case the brain-wide activation difference between both task conditions) and as we were interested in the unique control properties independent of traditional activation differences, we additionally controlled for the difference in the mean baseline energy in each analysis involving control energy. Baseline energy was defined as the absolute average difference in the unthresholded GLM parameter estimates over all regions between 0-back and 2-back conditions. As the same arguments apply to stability measures, we controlled for the respective baseline energy of each condition in each analysis involving stability. In all analyses involving polygenic scores for D1-/D2-expression, we further used the first 5 principal components from the PCA on the linkage-disequilibrium pruned set of autosomal SNPs to control for population stratification. Differences between control energy/stability of both conditions were assessed using repeated-measures ANOVA with condition as a within-subject factor. Drug effects were modeled using a repeated measures ANOVA with drug and condition as within-subject factors. As healthy control 11 and schizophrenia patients differed in one of the quantified DTI imaging quality parameters (tSNR, see Table S1), we also controlled for tSNR in each analysis involving both groups. 9. Control analyses 9.1. Null models of structural brain networks To study the impact of structural brain networks on control properties, we repeated the computation of control energy using a randomized null model of the individuals' structural brain networks that preserves the average weight, the strength distribution and the degree distribution. Null models were created using the null_model_und_sign function as implemented in the Brain Connectivity Toolbox (https://sites.google.com/site/bctnet/). In line with our expectation, control energy increased significantly for randomized networks in both groups (repeated measures ANOVA with null_model_vs_data and transition as within-subject factors, HC: main effect of null_model_vs_data, F(1,174) = 38.284, p < 0.001; SZ: F(1,20) = 3.561, p = 0.074), suggesting that human brain structural networks are in some form optimized to control brain state transitions, but schizophrenia patients tend to have less optimal, more random structural brain networks. 9.2. Null models of spatial activity patterns To study the impact of the spatial distribution of activity patterns on control properties, we repeated the computation of control energy and spatially randomized individuals' brain activation patterns. In line with our expectation, control energy increased significantly for randomized networks in both groups (repeated measures ANOVA with model_vs_data and transition as within-subject factors, HC: main effect of model_vs_data, F(1,174) = 6.995, p = 0.009; SZ: F(1,20) = 0.019, p < 0.0893), suggesting that the spatial distribution of brain activity patterns is important for minimizing control effort. 9.3. Robustness to choice of parcellation scheme To demonstrate the robustness of the results to our choice of parcellation scheme, we repeated our analysis using a recently published functionally defined atlas comprising a similar number of areas (31). Specifically, we used the "Gordon" template (31) consisting of 333 regions that are functionally derived from 12 resting-state connectivity analyses. Data were reprocessed using the same pipeline as for the main analysis and all parameters were kept identical in the subsequent analysis. Notably, we replicated all main results (see Table S4), indicating that our reported findings are robust to the choice of parcellation scheme. 9.4. Robustness to choice of edge definition To demonstrate the robustness of our results to our selection of connectivity measure, we repeated our analysis using the number of streamlines normalized by the respective size of the regions to construct structural connectivity matrices (15). All parameters were kept identical in the subsequent analysis. All main results could be replicated (see Table S4), indicating that our findings are robust to the choice of edge definition. 9.5. Impact of medication and duration of illness on control properties In patients, the potential relationship between control energy and stability, antipsychotic drug dose (expressed in chlorpromazine equivalents (CPZE), n=20), and clinical parameters (illness duration, illness severity as indexed by global functioning (GAF) and Positive and Negative Symptom Scale (PANSS)) were explored using Pearson correlation. Neither the control energy for the 0-back to 2-back transition nor the opposite transition or the stability of either state were significantly associated with CPZE (N = 20, 0- to 2- back: r = 0.078, p = 0.767; 2- to 0-back: r = 0.320, p = 0.210; 0- back stability: r = 0.150, p = 0.564; 2- back stability: r = 0.096, p = 0.713), with illness duration (N = 23, 0- to 2-back: r = 0.017, p = 0.937; 2- to 0-back: r = -0.226, p = 0.299; 0- back stability: r = 0.110, p = 0.644; 2- back stability: r = 0.281, p = 0.230), or with GAF (N = 24, 0- to 2-back: r = -0.086, p = 0.690; 2- to 0-back: r = -0.254, p = 0.230; 0- back stability: r = - 0.135, p = 0.570; 2- back stability: r = 0.066, p = 0.793). 9.6. Pharmacological validation using Risperidone To demonstrate the robustness of our pharmacological intervention of dopaminergic signaling, we additionally analyzed the data of the Risperidone condition in the same subjects. Risperidone also preferentially targets D2 receptors, but also affects D1, adrenergic, serotoninergic and histaminergic pathways. Using the same models and covariates as in the main analysis, we detected a trend-wise increase in control energy needed for both transitions (repeated measures ANOVA with drug and transition as within-person factors; main effect of drug: F(1,10) = 3.490, p = 0.091; drug-by-condition interaction: 13 F(1,10) = 0.238, p = 0.636; activity difference, drug order, and sex as covariates of no interest), but no effect on stability (F(1,8) = 0.105, p = 0.334; mean brain activity, sex, and drug order as covariates of no interest). Although these results showed only trend-wise significance, likely due to the lower D2-specificity of Risperidone, the detected pattern was conserved across drugs, validating the proposed underlying concepts. 9.7. Null results for gene score and imaging associations As mentioned in the main text, D1 receptor expression-related gene scores predicted stability of both states (0-back: b = 0.184, p = 0.034; 2-back: b = 0.242 p = 0.007), but not D2 receptor expression-related gene scores (0-back: b = 0.153, p = 0.109; 2-back: b = -0.01 p = 0.924). In turn, the control energy of both state transitions could be predicted by the D2 receptor expression-related score (0- to 2-back: b = -0.076, p = 0.037; and trending for 2- to 0-back: b = -0.134, p = 0.068), but not by the D1 receptor expression-related gene score (0- to 2-back: b = -0.037, p = 0.324; 2- to 0-back: b = -0.06, p = 0.418). 9.8. Suboptimal trajectories As mentioned in the main text, the variability in suboptimal trajectories was greater in schizophrenia (rm- ANOVA: main effect of group: F(1,98) = 4.789, p = 0.031, controlling for age, sex, DTI tSNR ,brain state energy difference). These results remained significant after additionally accounting for the stability of both states and for the control energy of both transitions (rm-ANOVA: main effect of group: F(1,95) = 11.2, p = 0.001). 14 10. Supplementary figures 10.1. Figure S1: N-back task design Design of the N-back task: Stimuli were presented in blocks of either 0-back (left) or 2-back (right) conditions. There was no additional control or resting condition. In the 0-back condition, subjects were instructed to press the button on the response box corresponding to the number currently displayed on the presentation screen. Here, the red numbers next to the screen images on the left indicate correct responses. In the 2-back condition, subjects were instructed to press the button on the response box corresponding to the number presented two stimuli before the number currently displayed on the presentation screen. Here, the red numbers next to the screen images on the right indicate the correct responses. Each condition block lasted 30 seconds and was repeated four times in an interleaved manner as shown on the bottom of the figure. 15 10.2. Figure S2: Network control theory concept and methods A summary of the methods to assess brain dynamics using network control theory. We used a multimodal atlas, and applied it to both diffusion tensor imaging data to obtain a structural connectome, and to functional magnetic resonance imaging data to obtain activation patterns during a 0-back motor and attention control task and during a 2-back working memory task. Finally, we use network control theory to explain transitions between 0-back and 2-back states based on the underlying structural connectome. 16 11. Supplementary tables 11.1. Table S1: Characteristics of the healthy control and patient samples Healthy controls (n = 178) 33.05 ± 10.98 13.66 ± 2.41 30.74 ± 3.84 n.a n.a n.a n.a. Demographic information Age (year) Sex (male / female) 93 / 85 Years of education Psychological assessments MWTB PANSS positive PANSS negative BDI Years of illness fMRI task performance Accuracy (%) Reaction time (ms) Imaging quality parameters fMRI: mean frame- wise displacement (mm) DTI: mean absolute root-mean-square displacement (mm) DTI: tSNR 1.27 ± 0.74 5.61 ± 0.45 0.15 ± 0.06 68.75 ± 19.33 80.10 ± 18.35 496.15 ± 279.50 589.21 ± 286.80 627.06 ± 306.99 65.54 ± 19.79 Matched controls (n = 80) 35.49 ± 10.55 46 / 34 13.65 ± 2.73 30.32 ± 4.83 n.a. n.a n.a. n.a. Schizophrenia patients (n = 24) 32.25 ± 10.33 18 / 6 11.68 ± 1.45 29.13 ±3.27 12.50 ± 6.76 15.17 ± 6.76 12.42 ± 7.71 10.22 ± 9.32 t or χ² value P value 1.32 2.39 2.72 1.11 - - - - 0.70 -0.6 0.188 0.122 0.008 0.272 - - - - 0.479 0.553 0.18 ± 0.07 0.20 ± 0.09 -1.11 0.270 1.37 ± 0.89 1.34 ± 0.59 0.18 0.860 5.52 ± 0.49 5.26 ± 0.49 2.21 0.028 Abbreviations: MWTB = Mehrfach Wortschatz Intelligenztest B, a German multiple-choice vocabulary intelligence test as a measure of premorbid IQ; PANSS = positive and negative symptom scale; BDI = Beck's depression inventory; DTI = diffusion tensor imaging; tSNR temporal signal-to-noise. 17 11.2. Table S2: Sample characteristics of the pharmacological study Healthy control (n = 16) t value P value Placebo Amisulpride Demographic information Age (year) Sex (male / female) fMRI task performance * 26.63 ± 5.34 8 / 8 Accuracy (%) 89.32 ± 10.00 85.68 ± 11.66 1.27 0.223 Reaction time (ms) 347.42 ± 104.26 365.40 ± 142.54 -0.582 0.569 Head motion parameters 0.124 ± 0.03 0.126 ± 0.04 -0.301 0.767 fMRI: Mean frame- wise displacement (mm) 18 11.3. Table S3: Statistical details for the main findings and replication analyses Glasser count Gordon FA p-val df F or t p-val df F or t p-val (stand beta) Result FA F or t (stand beta) df General properties of control stability 0 back > 2 back T02 > T20 stability 2-back -> accuracy stability 2-back -> RT Differential relation to D1 and D2 expression D1 -> stability 0 back D1 -> stability 2 back D2 -> stability 0 back D2 -> stability 2 back D2 -> T02 D2 -> T20 D1 -> T02 D1 -> T20 Drug effect on transition energy Pla vs. Ami Pla vs. Ris Drug effect on stability Pla vs. Ami Pla vs. Ris Schizophrenia stability 2-back: HC vs. SZ stability 0-back: HC vs. SZ T02: HC vs. SZ T20: HC vs. SZ 1,173 1,174 1,10 1,10 1,8 1,8 1,98 1,98 1,98 1,98 66.80 < 0.001 1,173 27.98 <0.001 1,174 -2.78 (0.274) -1.94 (-0.192) 0.006 0.054 2.18 (0.184) 2.78 (0.242) 1.629 (0.153) -0.095 (-0.010) -2.14 (-0.076) -1.87 (-0.134) -0.996 (-0.037) -0.817 (-0.06) 7.272 3.49 0.715 1.057 6.436 0.041 5.238 0.620 0.034 0.007 0.109 0.924 0.037 0.068 0.324 0.418 0.022 0.091 0.422 0.334 0.013 0.840 0.024 0.433 1,10 1,10 1,8 1,8 1,98 1,98 1,98 1,98 19 60.50 19.73 2.70 (0.252) -1.94 (-0.182) 2.212 (0.190) 2.978 (0.270) 1.963 (0.185) 0071 (0.008) -2.33 (-0.09) -1.83 (-0.134) -1.066 (-0.045) -0.650 (-0.049) 4.954 3.797 0.698 2.941 6.552 0.105 6.414 0.534 <0.001 1,173 <0.001 1,174 0.008 0.054 0.031 0.004 0.055 0.944 0.023 0.073 0.291 0.519 0.428 0.125 0.05 1,10 0.08 1,10 1,8 1,8 0.012 1,98 0.746 1,98 0.013 1,98 0.467 1,98 (stand beta) 56.846 21.706 2.58 (-0.239) -1.95 (-0.181) 3.10 (0.270) 3.39 (0.307) 0.146 (0.147) 0.174 (0.019) -2.07 (-0.07) -2.16 (-0.152) -0.832 (-0.029) -0.376 (-0.028) 8.839 3.014 0.013 0.358 4.951 0.327 5.070 0.275 <0.001 <0.001 0.011 0.053 0.003 0.001 0.149 0.862 0.043 0.036 0.414 0.709 0.014 0.113 0.913 0.566 0.028 0.569 0.027 0.601 Abbreviations: FA = structural edge weight defined as mean fraction anisotropy of a track connecting two regions; Count = structural edge weight defined as track count connecting two regions; Gordon = resting- state defined atlas with 333 regions; T02 = control energy for the transition 0 to 2-back; T20 = control energy for the transition 2 to 0-back; -> = predicts in a regression model; Q = modularity estimate; Pla = Placebo; Ami = Amisulpride; Ris = Risperidone; HC = healthy control; SZ = schizophrenia patients. 20 11.4. Table S4: List of brain regions included in the extended Glasser parcellation # Brain Region Label 1 Glasser_L_V1 2 Glasser_L_MST 3 Glasser_L_V6 4 Glasser_L_V2 5 Glasser_L_V3 6 Glasser_L_V4 7 Glasser_L_V8 8 Glasser_L_4 9 Glasser_L_3b 10 Glasser_L_FEF 11 Glasser_L_PEF 12 Glasser_L_55b 13 Glasser_L_V3A 14 Glasser_L_RSC 15 Glasser_L_POS2 16 Glasser_L_V7 17 Glasser_L_IPS1 18 Glasser_L_FFC 19 Glasser_L_V3B 20 Glasser_L_LO1 21 Glasser_L_LO2 22 Glasser_L_PIT 23 Glasser_L_MT 24 Glasser_L_A1 25 Glasser_L_PSL 26 Glasser_L_SFL 27 Glasser_L_PCV 28 Glasser_L_STV 29 Glasser_L_7Pm 30 Glasser_L_7m 31 Glasser_L_POS1 32 Glasser_L_23d 33 Glasser_L_v23ab 34 Glasser_L_d23ab 35 Glasser_L_31pv 36 Glasser_L_5m 37 Glasser_L_5mv 38 Glasser_L_23c 39 Glasser_L_5L 40 Glasser_L_24dd 41 Glasser_L_24dv 42 Glasser_L_7AL 43 Glasser_L_SCEF 44 Glasser_L_6ma 45 Glasser_L_7Am 46 Glasser_L_7PL 47 Glasser_L_7PC 48 Glasser_L_LIPv 49 Glasser_L_VIP 50 Glasser_L_MIP 51 Glasser_L_1 52 Glasser_L_2 53 Glasser_L_3a 54 Glasser_L_6d 55 Glasser_L_6mp 56 Glasser_L_6v 57 Glasser_L_p24pr 58 Glasser_L_33pr 59 Glasser_L_a24pr 60 Glasser_L_p32pr 61 Glasser_L_a24 62 Glasser_L_d32 63 Glasser_L_8BM 64 Glasser_L_p32 65 Glasser_L_10r 66 Glasser_L_47m 67 Glasser_L_8Av 68 Glasser_L_8Ad 69 Glasser_L_9m 70 Glasser_L_8BL 71 Glasser_L_9p 72 Glasser_L_10d 73 Glasser_L_8C 74 Glasser_L_44 75 Glasser_L_45 76 Glasser_L_47l 77 Glasser_L_a47r 78 Glasser_L_6r 79 Glasser_L_IFJa 80 Glasser_L_IFJp 81 Glasser_L_IFSp 82 Glasser_L_IFSa 83 Glasser_L_p9-46v 84 Glasser_L_46 85 Glasser_L_a9-46v 86 Glasser_L_9-46d 87 Glasser_L_9a 88 Glasser_L_10v 89 Glasser_L_a10p 90 Glasser_L_10pp 91 Glasser_L_11l 92 Glasser_L_13l 93 Glasser_L_OFC 94 Glasser_L_47s 95 Glasser_L_LIPd 96 Glasser_L_6a 97 Glasser_L_i6-8 98 Glasser_L_s6-8 99 Glasser_L_43 100 Glasser_L_OP4 101 Glasser_L_OP1 102 Glasser_L_OP2-3 103 Glasser_L_52 104 Glasser_L_RI 105 Glasser_L_PFcm 106 Glasser_L_PoI2 107 Glasser_L_TA2 108 Glasser_L_FOP4 109 Glasser_L_MI 110 Glasser_L_Pir 111 Glasser_L_AVI 112 Glasser_L_AAIC 113 Glasser_L_FOP1 114 Glasser_L_FOP3 115 Glasser_L_FOP2 116 Glasser_L_PFt 117 Glasser_L_AIP 118 Glasser_L_EC 119 Glasser_L_PreS 120 Glasser_L_H 121 Glasser_L_ProS 122 Glasser_L_PeEc 123 Glasser_L_STGa 124 Glasser_L_PBelt 125 Glasser_L_A5 126 Glasser_L_PHA1 127 Glasser_L_PHA3 128 Glasser_L_STSda 129 Glasser_L_STSdp 130 Glasser_L_STSvp 131 Glasser_L_TGd 132 Glasser_L_TE1a 133 Glasser_L_TE1p 134 Glasser_L_TE2a 135 Glasser_L_TF 136 Glasser_L_TE2p 137 Glasser_L_PHT 138 Glasser_L_PH 139 Glasser_L_TPOJ1 140 Glasser_L_TPOJ2 141 Glasser_L_TPOJ3 142 Glasser_L_DVT 143 Glasser_L_PGp 144 Glasser_L_IP2 145 Glasser_L_IP1 146 Glasser_L_IP0 147 Glasser_L_PFop 148 Glasser_L_PF 149 Glasser_L_PFm 21 150 Glasser_L_PGi 151 Glasser_L_PGs 152 Glasser_L_V6A 153 Glasser_L_VMV1 154 Glasser_L_VMV3 155 Glasser_L_PHA2 156 Glasser_L_V4t 157 Glasser_L_FST 158 Glasser_L_V3CD 159 Glasser_L_LO3 160 Glasser_L_VMV2 161 Glasser_L_31pd 162 Glasser_L_31a 163 Glasser_L_VVC 164 Glasser_L_25 165 Glasser_L_s32 166 Glasser_L_pOFC 167 Glasser_L_PoI1 168 Glasser_L_Ig 169 Glasser_L_FOP5 170 Glasser_L_p10p 171 Glasser_L_p47r 172 Glasser_L_TGv 173 Glasser_L_MBelt 174 Glasser_L_LBelt 175 Glasser_L_A4 176 Glasser_L_STSva 177 Glasser_L_TE1m 178 Glasser_L_PI 179 Glasser_L_a32pr 180 Glasser_L_p24 181 Glasser_R_V1 182 Glasser_R_MST 183 Glasser_R_V6 184 Glasser_R_V2 185 Glasser_R_V3 186 Glasser_R_V4 187 Glasser_R_V8 188 Glasser_R_4 189 Glasser_R_3b 190 Glasser_R_FEF 191 Glasser_R_PEF 192 Glasser_R_55b 193 Glasser_R_V3A 194 Glasser_R_RSC 195 Glasser_R_POS2 196 Glasser_R_V7 197 Glasser_R_IPS1 198 Glasser_R_FFC 199 Glasser_R_V3B 200 Glasser_R_LO1 201 Glasser_R_LO2 202 Glasser_R_PIT 203 Glasser_R_MT 204 Glasser_R_A1 205 Glasser_R_PSL 206 Glasser_R_SFL 207 Glasser_R_PCV 208 Glasser_R_STV 209 Glasser_R_7Pm 210 Glasser_R_7m 211 Glasser_R_POS1 212 Glasser_R_23d 213 Glasser_R_v23ab 214 Glasser_R_d23ab 215 Glasser_R_31pv 216 Glasser_R_5m 217 Glasser_R_5mv 218 Glasser_R_23c 219 Glasser_R_5L 220 Glasser_R_24dd 221 Glasser_R_24dv 222 Glasser_R_7AL 223 Glasser_R_SCEF 224 Glasser_R_6ma 225 Glasser_R_7Am 226 Glasser_R_7PL 227 Glasser_R_7PC 228 Glasser_R_LIPv 229 Glasser_R_VIP 230 Glasser_R_MIP 231 Glasser_R_1 232 Glasser_R_2 233 Glasser_R_3a 234 Glasser_R_6d 235 Glasser_R_6mp 236 Glasser_R_6v 237 Glasser_R_p24pr 238 Glasser_R_33pr 239 Glasser_R_a24pr 240 Glasser_R_p32pr 241 Glasser_R_a24 242 Glasser_R_d32 243 Glasser_R_8BM 244 Glasser_R_p32 245 Glasser_R_10r 246 Glasser_R_47m 247 Glasser_R_8Av 248 Glasser_R_8Ad 249 Glasser_R_9m 22 250 Glasser_R_8BL 251 Glasser_R_9p 252 Glasser_R_10d 253 Glasser_R_8C 254 Glasser_R_44 255 Glasser_R_45 256 Glasser_R_47l 257 Glasser_R_a47r 258 Glasser_R_6r 259 Glasser_R_IFJa 260 Glasser_R_IFJp 261 Glasser_R_IFSp 262 Glasser_R_IFSa 263 Glasser_R_p9-46v 264 Glasser_R_46 265 Glasser_R_a9-46v 266 Glasser_R_9-46d 267 Glasser_R_9a 268 Glasser_R_10v 269 Glasser_R_a10p 270 Glasser_R_10pp 271 Glasser_R_11l 272 Glasser_R_13l 273 Glasser_R_OFC 274 Glasser_R_47s 275 Glasser_R_LIPd 276 Glasser_R_6a 277 Glasser_R_i6-8 278 Glasser_R_s6-8 279 Glasser_R_43 280 Glasser_R_OP4 281 Glasser_R_OP1 282 Glasser_R_OP2-3 283 Glasser_R_52 284 Glasser_R_RI 285 Glasser_R_PFcm 286 Glasser_R_PoI2 287 Glasser_R_TA2 288 Glasser_R_FOP4 289 Glasser_R_MI 290 Glasser_R_Pir 291 Glasser_R_AVI 292 Glasser_R_AAIC 293 Glasser_R_FOP1 294 Glasser_R_FOP3 295 Glasser_R_FOP2 296 Glasser_R_PFt 297 Glasser_R_AIP 298 Glasser_R_EC 299 Glasser_R_PreS 300 Glasser_R_H 301 Glasser_R_ProS 302 Glasser_R_PeEc 303 Glasser_R_STGa 304 Glasser_R_PBelt 305 Glasser_R_A5 306 Glasser_R_PHA1 307 Glasser_R_PHA3 308 Glasser_R_STSda 309 Glasser_R_STSdp 310 Glasser_R_STSvp 311 Glasser_R_TGd 312 Glasser_R_TE1a 313 Glasser_R_TE1p 314 Glasser_R_TE2a 315 Glasser_R_TF 316 Glasser_R_TE2p 317 Glasser_R_PHT 318 Glasser_R_PH 319 Glasser_R_TPOJ1 320 Glasser_R_TPOJ2 321 Glasser_R_TPOJ3 322 Glasser_R_DVT 323 Glasser_R_PGp 324 Glasser_R_IP2 325 Glasser_R_IP1 326 Glasser_R_IP0 327 Glasser_R_PFop 328 Glasser_R_PF 329 Glasser_R_PFm 330 Glasser_R_PGi 331 Glasser_R_PGs 332 Glasser_R_V6A 333 Glasser_R_VMV1 334 Glasser_R_VMV3 335 Glasser_R_PHA2 336 Glasser_R_V4t 337 Glasser_R_FST 338 Glasser_R_V3CD 339 Glasser_R_LO3 340 Glasser_R_VMV2 341 Glasser_R_31pd 342 Glasser_R_31a 343 Glasser_R_VVC 344 Glasser_R_25 345 Glasser_R_s32 346 Glasser_R_pOFC 347 Glasser_R_PoI1 348 Glasser_R_Ig 349 Glasser_R_FOP5 350 Glasser_R_p10p 351 Glasser_R_p47r 352 Glasser_R_TGv 353 Glasser_R_MBelt 354 Glasser_R_LBelt 355 Glasser_R_A4 356 Glasser_R_STSva 357 Glasser_R_TE1m 358 Glasser_R_PI 359 Glasser_R_a32pr 360 Glasser_R_p24 504 HO_Left_Thalamus 505 HO_Left_Caudate 506 HO_Left_Putamen 507 HO_Left_Pallidum 509 HO_Left_Hippocampus 510 HO_Left_Amygdala 511 HO_Left_Accumbens 515 HO_Right_Thalamus 516 HO_Right_Caudate 517 HO_Right_Putamen 518 HO_Right_Pallidum 519 HO_Right_Hippocampus 520 HO_Right_Amygdala 521 HO_Right_Accumbens 23 12. Supplementary references 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. Tombaugh TN (2004) Trail Making Test A and B: normative data stratified by age and education. Archives of clinical neuropsychology 19(2):203-214. Lehrl S, Triebig G, & Fischer B (1995) Multiple choice vocabulary test MWT as a valid and short test to estimate premorbid intelligence. Acta Neurologica Scandinavica 91(5):335-345. Kay SR, Fiszbein A, & Opler LA (1987) The positive and negative syndrome scale (PANSS) for schizophrenia. Schizophrenia bulletin 13(2):261-276. Beck AT, Ward CH, Mendelson M, Mock J, & Erbaugh J (1961) An inventory for measuring depression. Arch Gen Psychiatry 4:561-571. Braun U, et al. (2015) Dynamic reconfiguration of frontal brain networks during executive cognition in humans. Proc Natl Acad Sci U S A 112(37):11678-11683. Plichta MM, et al. (2012) Test-retest reliability of evoked BOLD signals from a cognitive-emotive fMRI test battery. Neuroimage 60(3):1746-1758. Callicott JH, et al. (1999) Physiological characteristics of capacity constraints in working memory as revealed by functional MRI. Cerebral Cortex 9(1):20-26. Glasser MF, et al. (2016) A multi-modal parcellation of human cerebral cortex. Nature 536(7615):171- 178. Smith SM, et al. (2004) Advances in functional and structural MR image analysis and implementation as FSL. Neuroimage 23 Suppl 1:S208-219. Smith SM (2002) Fast robust automated brain extraction. Hum Brain Mapp 17(3):143-155. Yeh FC, Verstynen TD, Wang Y, Fernandez-Miranda JC, & Tseng WY (2013) Deterministic diffusion fiber tracking improved by quantitative anisotropy. PLoS One 8(11):e80713. Baum GL, et al. (2018) The impact of in-scanner head motion on structural connectivity derived from diffusion MRI. Neuroimage 173:275-286. Chung MH, et al. (2018) Individual differences in rate of acquiring stable neural representations of tasks in fMRI. PLoS One 13(11):e0207352. Braun U, et al. (2016) Dynamic brain network reconfiguration as a potential schizophrenia genetic risk mechanism modulated by NMDA receptor function. Proc Natl Acad Sci U S A 113(44):12568-12573. Betzel RF, Gu S, Medaglia JD, Pasqualetti F, & Bassett DS (2016) Optimally controlling the human connectome: the role of network topology. Sci Rep 6:30770. Gu S, et al. (2017) Optimal trajectories of brain state transitions. Neuroimage 148:305-317. Kim JZ, et al. (2018) Role of Graph Architecture in Controlling Dynamical Networks with Applications to Neural Systems. Nat Phys 14:91-98. Cornblath EJ, et al. (2018) Context-dependent architecture of brain state dynamics is explained by white matter connectivity and theories of network control. arXiv preprint arXiv:1809.02849. Howie B, Fuchsberger C, Stephens M, Marchini J, & Abecasis GR (2012) Fast and accurate genotype imputation in genome-wide association studies through pre-phasing. Nat Genet 44(8):955-959. Howie BN, Donnelly P, & Marchini J (2009) A flexible and accurate genotype imputation method for the next generation of genome-wide association studies. PLoS genetics 5(6):e1000529. Delaneau O, Zagury JF, & Marchini J (2013) Improved whole-chromosome phasing for disease and population genetic studies. Nature methods 10(1):5-6. Fazio L, et al. (2018) Transcriptomic context of DRD1 is associated with prefrontal activity and behavior during working memory. Proc Natl Acad Sci U S A 115(21):5582-5587. Pergola G, et al. (2017) DRD2 co-expression network and a related polygenic index predict imaging, behavioral and clinical phenotypes linked to schizophrenia. Transl Psychiatry 7(1):e1006. Selvaggi P, et al. (2019) Genetic Variation of a DRD2 Co-expression Network is Associated with Changes in Prefrontal Function After D2 Receptors Stimulation. Cereb Cortex 29(3):1162-1173. Fazio L, et al. (2018) Transcriptomic context of DRD1 is associated with prefrontal activity and behavior during working memory. Proceedings of the National Academy of Sciences of the United States of America. Selvaggi P, et al. (2018) Genetic Variation of a DRD2 Co-expression Network is Associated with Changes in Prefrontal Function After D2 Receptors Stimulation. Cerebral cortex. 24 27. 28. 29. 30. 31. Pergola G, et al. (2018) Prefrontal co-expression of schizophrenia risk genes is associated with treatment response in patients. bioRxiv. Zhang B & Horvath S (2005) A general framework for weighted gene co-expression network analysis. Statistical applications in genetics and molecular biology 4:Article17. Colantuoni C, et al. (2011) Temporal dynamics and genetic control of transcription in the human prefrontal cortex. Nature 478(7370):519-523. Pergola G, et al. (2016) Combined effect of genetic variants in the GluN2B coding gene (GRIN2B) on prefrontal function during working memory performance. Psychological medicine 46(6):1135-1150. Gordon EM, et al. (2016) Generation and Evaluation of a Cortical Area Parcellation from Resting-State Correlations. Cereb Cortex 26(1):288-303. 25
1807.02126
1
1807
2018-07-05T18:03:19
Cognitive chimera states in human brain networks
[ "q-bio.NC", "math.DS", "physics.data-an" ]
The human brain is a complex dynamical system that gives rise to cognition through spatiotemporal patterns of coherent and incoherent activity between brain regions. As different regions dynamically interact to perform cognitive tasks, variable patterns of partial synchrony can be observed, forming chimera states. We propose that the emergence of such states plays a fundamental role in the cognitive organization of the brain, and present a novel cognitively-informed, chimera-based framework to explore how large-scale brain architecture affects brain dynamics and function. Using personalized brain network models, we systematically study how regional brain stimulation produces different patterns of synchronization across predefined cognitive systems. We then analyze these emergent patterns within our novel framework to understand the impact of subject-specific and region-specific structural variability on brain dynamics. Our results suggest a classification of cognitive systems into four groups with differing levels of subject and regional variability that reflect their different functional roles.
q-bio.NC
q-bio
Cognitive chimera states in human brain networks July 9, 2018 Kanika Bansal1,2,3,*, Javier O. Garcia1,4, Steven H. Tompson1,4, Timothy Verstynen5, Jean M. Vettel1,4,6, Sarah F. Muldoon3,7,* 1 Human Sciences, US Army Research Laboratory, Aberdeen Proving Ground, MD 21005, USA 2 Department of Biomedical Engineering, Columbia University, New York, NY 10027, USA 3 Mathematics Department, University at Buffalo – SUNY, Buffalo, NY 14260, USA 4 Department of Biomedical Engineering, University of Pennsylvania, Philadelphia, PA 19104, USA 5 Department of Psychology, Carnegie Mellon University, Pittsburgh, PA 15213, USA 6 Department of Psychological and Brain Sciences, University of California, Santa Barbara, CA 93106 , USA 7 CDSE Program and Neuroscience Program, University at Buffalo – SUNY, Buffalo, NY 14260, USA *Correspondence should be addressed to KB ([email protected]) and SFM ([email protected]). 8 1 0 2 l u J 5 ] . C N o i b - q [ 1 v 6 2 1 2 0 . 7 0 8 1 : v i X r a 1 Abstract The human brain is a complex dynamical system that gives rise to cognition through spatiotemporal patterns of coherent and incoherent activity between brain regions. As different regions dynamically interact to perform cognitive tasks, variable patterns of partial synchrony can be observed, forming chimera states. We propose that the emergence of such states plays a fundamental role in the cognitive organization of the brain, and present a novel cognitively-informed, chimera-based framework to explore how large-scale brain architecture affects brain dynamics and function. Using personalized brain network models, we systematically study how regional brain stimulation produces different patterns of synchro- nization across predefined cognitive systems. We then analyze these emergent patterns within our novel framework to understand the impact of subject-specific and region-specific structural variability on brain dynamics. Our results suggest a classification of cognitive systems into four groups with differing levels of subject and regional variability that reflect their different functional roles. 2 Introduction Rhythmic behavior is ubiquitous in complex systems, and a diverse set of research has examined how in- teracting system elements come together to form synchronized, coherent behavior across domains that span biological [1, 2], social [3], and engineered [4] settings. However, the emergence of complete system-wide synchronization might not always provide the best description of system dynamics. In many systems, states of partial synchrony have been observed, where a system organizes in separate domains of synchronized elements [5, 6, 7]. This is particularly true in the human brain, where patterns of neurophysiological activity evolve rapidly, showing transient domains of synchronization across subsets of brain regions [8, 9, 5]. In the last decade, the rise of network neuroscience approaches [10, 11] have demonstrated a foundational role for partial synchrony among separate cognitive sub-networks, where the underlying architecture of the brain ensures efficient integration of sensory input with stored knowledge while also segregating task-irrelevant in- formation to support cognition [12, 13]. However, the fundamental principles and constraints that subserve the intricate timing and specificity of these time-evolving patterns of synchrony are not well understood [14]. The dynamical systems framework of chimera states offers a powerful tool to study the evolution of coherent and incoherent dynamics in oscillating systems such as the brain. A chimera state emerges when a system of oscillators evolves into two subsets of mutually coherent and incoherent populations [15, 16]. Although chimera states represent a natural link between coherent and incoherent dynamics [6, 17, 18], they were initially only explored analytically and their relationship to physical systems was unknown. In fact, it wasn't until almost a decade after their theoretical discovery that chimera states were demonstrated experimentally in opto-electronic [19], and chemical [20] oscillators, finally establishing their connection with real-world systems. Recently, the chimera framework has been extended to include states of multi-chimera [21, 22], traveling chimera [23], amplitude chimera, and chimera death [24], demonstrating the versatility of the framework to describe critical system dynamics. Because of its natural ability to describe patterns of partial synchronization, the chimera framework has an intuitive utility for augmenting our understanding of the brain. Patterns of synchronization between cognitive systems are thought to form the basis of cognition, and the interplay of synchrony among subsets of brain regions has been shown to underlie sleep-wake states [25], variability in task performance [12], and the continuum between healthy and disease states [26, 27]. Indeed, recent work has speculated that similarities exist between chimera states and brain dynamics during unihemispheric sleep [28], the transition to a seizure state in epilepsy [29], and coordinated finger tapping exercises [30]. Fundamentally, these dynamics are the result of complex interactions between neuronal populations and are often modeled using networks of coupled oscillators [31]. As a result, despite the intuitive similarities between chimera states and brain dynamics, much of the work relating chimera dynamics to neuroscience thus far has focused on understanding chimera states at the level of neuronal networks, using mathematical modeling of networks of individual neurons with fewer elements and/or simplified connection topologies [21, 22, 32, 33]. Only recently have neuronal models been used to examine the possibility of the emergence of chimera-like states within large-scale brain networks derived from two well-characterized animal brains – C. elegans [34] and the cat cortex [35]. However, even in these instances, the network connectivities were modified for simplicity. Thus, there remains a gap between studies of chimera states and applications to large-scale functional patterns of brain activity thought to underlie cognition. This largely reflects the computational complexity of modeling whole-brain dynamics and identifying an informative, yet simplified, model of cognitive processing. Here, we bridge this gap by presenting a novel cognitively-informed, chimera-based framework combined with in silico experiments, where we leverage the existence of a core set of predefined cognitive systems that constitute the functional organization of the brain [36, 37]. Consequently, our novel framework keeps the computational complexity of the analysis to a minimum while providing the unique ability to connect chimera states as an underlying basis for cognition. Using personalized brain network models (BNMs), we study cognitive system-level patterns of synchrony that emerge across 76 brain regions within 9 cognitive 3 systems as the result of regional brain stimulation. Our analysis focuses on how brain architecture relates to the frequency and types of dynamical patterns produced after stimulation. More specifically, we aim to answer two questions: (i) do patterns of synchronization observed for each cognitive system depend on what region was stimulated (region-specific effects), and (ii) does structural variability between participants decrease the consistency of patterns observed for each cognitive system (subject-specific effects)? From our in silico experiments, we observe different patterns of synchronization that can be classified into three dynamical states: (i) a coherent state of global synchrony, (ii) a chimera state with coexisting domains of synchrony and de-synchrony, and (iii) a metastable state with an absence of any large-scale stable synchrony. Our results demonstrate rich diversity in the states produced across all nine cognitive systems, including variability in patterns based on both region-specific and subject-specific structural variability. Critically, all nine cognitive systems give rise to chimera states and this likely reflects the foundational role that partial synchrony serves in large-scale brain function. Neuronal dynamics must concurrently segregate specialized processing while integrating localized functions for coordinated, cohesive cognitive performance. Our novel chimera-based framework provides an avenue to study how dynamical states give rise to variability in cognitive performance, providing the first approach that can uncover the link between chimera states and cognitive system functions that subserve human behavior. Results We first build subject-specific BNMs using anatomical connectivity derived from diffusion spectrum imaging data of 30 healthy individuals [38]. As illustrated in Fig. 1a, we parcellate the brain into 76 regions (network nodes) and define weighted network edges based on structural connectivity between brain regions. The dynamics of each brain region are modeled using Wilson-Cowan oscillators (WCOs), a biologically inspired, nonlinear mean-field model of a small population of neurons [39]. The coupling between regions is derived based on the unique structural connectivity of each individual (Fig. S1). When multiple WCOs are coupled, the resulting patterns of synchronization are highly dependent upon the topology of the coupling, ensuring that these BNMs are highly sensitive to individual variability in the underlying anatomical connectivity [36, 40, 41]. As illustrated in Fig. 1b and Fig. S2, we then perform in silico experiments by systematically applying computational regional (nodal) stimulation to the BNM and assessing the resulting patterns of synchronization that emerge based on subject-specific and region-specific variation in structural connectivity. Classification of brain states using cognitive systems. Traditionally, one would measure system-wide synchronization by calculating a measure such as the Ku- ramoto order parameter within the entire network of oscillators [42]. However, in order to study how the stimulation of different brain regions drives brain function, we instead focus on the relationship between patterns of synchronization between large-scale cognitive systems. Thus, based on previous research [36, 37], we assigned each of the 76 brain regions into one of nine cognitive systems (Fig. S3). Each cognitive system is defined by regions that coactivate in support of a generalized class of cognitive functions. This delin- eation includes several sensory-motor related systems, including auditory (Aud), visual (V), and motor and somatosensory (MS) systems, as well as the ventral temporal association system (VT) that encapsulates regions involved in knowledge representation. Several of the systems are involved in functional roles that are generic across cognitive performance, including the attention system (Att), the medial default mode sys- tem (mDM), and two systems associated with cognitive control, cingulo-opercular (CP) and frontoparietal (FP) systems. Finally, we include the subcortical system (SC) that consists of the regions responsible for autonomic and primal functions. In each in silico experiment, we stimulate a brain region in each subject-specific BNM and then compute cognitive-system-level synchronization. We calculate a cognitive-system-based Kuramoto order parameter 4 Figure 1: Design of the in silico experiments. (a) We construct personalized brain network models by estimating white matter anatomical connectivity of the brain using diffusion spectrum imaging. This con- nectivity is combined with a brain parcellation scheme with 76 cortical and subcortical regions to obtain large scale connectivity map of the regional brain volume. These regions constitute the nodes of the structural brain network, whose dynamics are simulated by nonlinear Wilson-Cowan oscillators, coupled through the structural connectivity map of a given individual (see Methods). (b) In the resulting data-driven models of the spatiotemporal dynamics of the brain, each brain region is systematically stimulated across a cohort of 30 subjects. The spread of the stimulation is measured through synchronization within the brain network. 5 ρci,cj that measures the amount of synchrony among all oscillators (regions) within two cognitive systems ci and cj, and obtain a cognitive-system-based synchronization matrix as shown in Fig. 2a. In this synchro- nization matrix we define two cognitive systems, ci and cj, to be synchronized if ρci,cj exceeds a threshold value (ρT h). In Fig. 2a we chose ρT h = 0.8 to define three dynamical states observed in this study: (i) a coherent state, where all cognitive systems are synchronized; (ii) a cognitive chimera state, where some cognitive systems form a synchronized cluster (yellow) while the other systems remain incoherent (blue); and (iii) a metastable state, where no stable synchrony between cognitive systems is observed. Next, we compare our cognitive-system-based analysis with two traditional measures of synchronization: (1) the classical Kuramoto order parameter [42] calculated across all 76 oscillators (regions) that captures the level of global synchrony in the network, and (2) the chimera-index [43, 34] that describes the closeness of the state with an ideal chimera state (see Methods). In Fig. 2b, we show how the three dynamical states (co- herent, chimera, and metastable) observed after stimulation of different brain regions in different individuals are distributed in the global synchrony and chimera-index parameter space. We observe a clear separation of these states, and as expected, global synchrony decreases from the coherent to chimera to metastable state. Thus, a cognitively-informed, systems-based classification of dynamical states is comparable to the traditional measures of estimating synchrony within a network. It is also robust across a range of threshold values (ρT h) as shown in Figures S4 and S5. We also examine how these two traditional metrics of synchronization relate to the connectivity properties of the node (region) itself. As seen in Fig. 2c, the level of global synchrony is positively correlated with the degree of the region being stimulated (r = 0.81; p = 1.8 × 10−124). Stimulation of a network hub (highly connected brain region) is therefore more likely to produce a coherent state, while stimulation of a non-hub is more likely to result in either a chimera or metastable state. Interestingly, Fig. 2d reveals that the chimera-index shows a weaker and negative correlation with degree (r = −0.57; p = 7.4 × 10−195). This relationship indicates the ability of moderately connected brain regions to produce a variety of spatially distinct synchronization patterns as a result of stimulation. These results not only demonstrate that chimera states emerge among large-scale cognitive systems but also reveal that the variable structural connectivity of regions within cognitive systems can drive the whole brain into diverse synchronization patterns. Variable brain states emerge from stimulation of different brain regions. Given that the type of brain state that emerges as result of stimulation is related to the local connectivity of the region, we next asked if there was also a spatial relationship between the location of the stimulated region and the type of dynamical state produced. In Fig. 3, brain regions are depicted as an orb, and their sizes denote the normalized occurrence of a given state, that they produce upon stimulation, calculated across subjects. Regions that produce coherent states (network hubs) are distributed more closely to the center of the brain (3a), while regions that produce the opposite extreme, metastable states, are distributed along the periphery of the brain (3c). Interestingly, regions that produce chimera states are located between the two (3b). This asymmetry in the distribution of states highlights the differences in the structural organization of the brain. Regions in the center of the brain that produce a coherent state (e.g., subcortical regions such as the hippocampus, thalamus) can play a global cognitive role and facilitate communication between spatially separated brain regions. Conversely, regions located in the periphery produce metastable states consistent with the notion that local and/or specialized computations take place in the cortex and could reflect the fact that the structure of the human brain has evolved to produce complex cognitive abilities [44]. Due to the diversity in the distribution of dynamical states across spatially distributed regions, we inves- tigate the relative contribution of the nine cognitive systems in producing each dynamical state after regional stimulation. As shown in Fig. 4a, coherent states are produced predominantly by regional stimulation within subcortical and medial default mode systems, reflecting that many of their constituent regions are network hubs. 6 Figure 2: Emergence of dynamical states within a cognitively-informed framework. (a) Cognitive system- level synchronization matrices whose entries denote the extent of synchronization between cognitive systems for coherent, chimera and metastable states. (b) The global synchronization and chimera-index of the system after stimulation of each brain region across all subjects. Within this traditional framework, our cognitively- defined states can be identified to have distinguishable characteristics. A coherent state shows high global synchronization value and low chimera-index (red). Both chimera (yellow) and metastable (blue) states show lower global synchronization. A chimera state can have either a higher global synchronization or a higher chimera-index than metastable states. (c)-(d) Origin of these states follows the connectivity of the stimulated brain region. (c) The global synchronization measure is positively correlated with the weighted degree of brain regions (r = 0.81, p = 1.8 × 10−124), indicating that the network hubs are more likely to produce a coherent state. (d) The Chimera-index is weakly and negatively correlated with the weighted degree (r = −0.57, p = 7.4 × 10−195), indicating that stimulation of lower degree nodes is more likely to produce an ideal chimera state with half of the population synchronized and the other half de-synchronized. 7 Figure 3: Distributed origin of dynamical states. Separate depictions of brain network regions (nodes) that produce each of the three dynamical states: (a) coherent, (b) chimera, and (c) metastable. Both axial and coronal views of the brain are presented to visualize the anatomical location of the node in left (LH) and right (RH) hemispheres. Node size represents the normalized occurrence of a given state upon stimulation of the node across all the subjects. Variability in node sizes imply that different states can emerge upon the stimulation of a single node across individuals. A coherent state is significantly more likely to be originated from nodes within the center of the brain. (b) Chimera states are likely to be originated by the stimulation of nodes that are relatively equally distributed within the brain; however, these nodes may vary in the spatial patterns of synchronization that they produce. (c) A metastable state is more likely to originate from the stimulation of peripheral brain regions. 8 Figure 4: Contribution of cognitive systems to dynamical states. (a) Coherent states are likely to result when nodes within medial default mode and subcortical systems are stimulated. (b) Chimera states emerge upon the stimulation of nodes that are equally distributed across all the cognitive systems. (c) Metastable states frequently occur after stimulation of nodes within auditory, cingulo-opercular, frontoparietal, and ventral temporal association systems. This distribution indicates the dominance of a particular type of cognitive role within the nodes of different cognitive systems. 9 There is also system specificity for metastable states which are preferentially produced by four systems (Fig. 4c). Two of those systems: cingulo-opercular and forntoparietal, are associated with cognitive control and proposed to be complementary systems that may often need to process task-relevant information concur- rently. Their dominance in producing metastable states likely reflects the fact that these systems can work in seclusion without co-activating a large part of the brain (facilitating parallel processing) and that they are flexible and not constrained by their structural connectivity. Both the auditory and the ventral temporal association systems also contribute significantly to metastable states. These two systems are both predomi- nantly located in the temporal lobe, an area of the brain associated with knowledge representation, so their functional roles may also frequently require working in seclusion in support of higher order perception. Although coherent and metastable states are dominantly produced by specific cognitive systems, all nine systems give rise to chimera states (Fig. 4b). In fact, chimera states are more likely to occur than either coherent or metastable states (Figures 3b and 2b) and encompass a variety of different spatial patterns of coexisting coherent and incoherent behavior. This likely reflects the foundational role of partial synchrony in large-scale brain function. Cognitive tasks constantly require the intricate balance between segregated and integrated neural processing [12]. The robust occurrence of chimera states following stimulation to each of the nine cognitive systems reflects the complexity and flexibility of the brains underlying architecture to support diverse processing requirements. Structural variability influences observed dynamic states. We next characterize the spatial patterns that comprise the chimera states to understand which cognitive systems are synchronized and de-synchronized following stimulation to each region. The results are organized by stimulation of brain regions within a cognitive system in Fig. 5a-i, where each row is a possible pattern of synchronization within the nine cognitive systems. The rows are organized by the frequency with which the pattern was observed (listed to the right of the row) after stimulation to all regions within that system and across all individuals in the study. For each pattern, systems that are part of the synchronized population after stimulation are shown in orange, and systems that are part of the de-synchronized population are shown in white. Consequently, coherent states are demarcated by a fully orange row, metastable states by a fully white row, and chimera states by a mixed pattern of coloring. For each cognitive system (Fig. 5a-i), we present the prevalent patterns observed and these results illustrate what systems are likely to synchronize after stimulation. Aligned with their complementary roles in cognitive control, the cingulo-opercular and frontoparietal systems continue to show similarity in their patterns of synchronization. The dominant pattern after stimu- lation to regions in either system is a metastable state, occurring 45% for cingulo-opercular system (Fig. 5c) and 50% for the frontoparietal system (Fig. 5d). Similarly, the auditory system also produces a metastable state 50% of the time (Fig. 5b). For all three of these systems, the second most frequent state is the opposite extreme, a coherent state (20% for auditory and 11% for cingulo-opercular) or nearly coherent state (7% for frontoparietal). Thus, these three systems show diversity in the types of dynamical states that they are capable of producing: sometimes stimulation of the system produces a metastable (segregated) state, while other times, stimulation of the system drives the brain to a coherent (integrated) state. The ventral temporal association system also produces a metastable state as its most prevalent pattern (Fig. 5h), but unlike in the three previously discussed systems, this state is produced less frequently (16%), and the system also produces a much larger variety of prevalent patterns of synchronization (10 unique patterns). The only other system to show this high level of diversity in its produced patterns is the motor and somatosensory system (10 patterns, Fig. 5f). In both systems, multiple patterns of chimera states are observed. This likely reflects the ubiquitous need for neural processing related to both action coordination (motor and somatosensory) and higher order perception (ventral temporal association) to be integrated with the processing occurring in other systems within the brain [45, 46]. 10 Figure 5: Patterns of synchronization and cognitive chimera states. (a)-(i) Prevalent patterns (with an occurrence of at least 3%) that emerge as the regions within different cognitive systems are stimulated across all subjects. Each panel represents stimulation of regions within a particular cognitive system. Each row represents one pattern of synchronization, and each column represents the state of a cognitive system. Cognitive systems that belong to the synchronized population are colored orange and cognitive systems that remain desynchronized are colored white. Thus, a fully orange or white row represents a coherent or metastable state respectively. Chimera states show different patterns of coloring depending on the cognitive systems that are recruited to the synchronized group. Different rows of patterns are stacked based on their relative occurrences (mentioned on the right side). To summarize the observed patterns, (j) shows the probability with which different cognitive systems can be synchronized when the regions within a given system are stimulated across subjects (shown along the vertical axis). 11 A coherent state occurs most frequently for the attention (36%, Fig. 5a), default mode (33%, Fig. 5e), and subcortical systems (43%, Fig. 5g). The visual system is also similar, though the coherent state is the second most prevalent (19%) with a nearly coherent state (all but auditory) as its dominant pattern (22%). Overall, these four systems are less dynamically diverse since their stimulation largely results in chimera states with high synchrony. All of these systems serve fundamental functional roles to rapidly respond to the external environment, and their dynamical patterns reflect this need to efficiently integrate this information with other cognitive systems. For all systems, the most common state following regional stimulation is a chimera state with high synchrony, emphasizing the importance of partial synchrony for all of the diverse functional roles provided by large-scale cognitive systems. Visual inspection of the patterns suggests that the systems most likely to belong to the desynchronized population of a chimera state are the three systems that predominantly produce a metastable state (auditory, cingulo-opercular, and frontoparietal systems). This effect is quantified in the synchronization probability plot (Fig. 5j). Systems with a high probability of synchronization have dark colors, while systems that are unlikely to be synchronized as a result of stimulation to a specific cognitive system are shown in light colors. While the auditory, cingulo-opercular, and frontoparietal systems systems are all unlikely to synchronize with other systems, the attention system, in contrast, is highly likely to be part of the synchronized population following stimulation to any system. We also observe that the stimulation of a region within a particular system does not necessarily induce synchronization within that system, which is particularly the case with the auditory, cingulo-opercular, and frontoparietal systems. Collectively, these results reveal the power of this approach to characterize large-scale system effects after regional stimulation. In Fig. S6, we further describe how different nodes in a system contribute to observed patterns by spatially mapping the probability of a given pattern onto the brain. Taken together, these results highlight that stimulation within each system gives rise to multiple patterns, similar patterns can emerge from spatially different regions, and within a system, there can be a special distribution of states across brain regions. This likely arises from individual differences in the structural connectivity between the participants in the study or differences in the structural connectivity of the regions themselves within each system (or combination of the two). Consequently, we introduce a new metric to assess the contribution of subject-specific and region-specific variability on the observed patterns. Dissociation of subject-specific and region-specific variability In our final analysis, we compute a measure called tenacity, which we defined to quantify the level of similarity between a set of observed patterns (see Methods), to assess how structural variability, either between subjects or between regions, influenced the patterns observed in Fig. 5. To differentiate these two potential sources of variability, we separately compute a subject tenacity and a region tenacity score (see Methods). When tenacity is calculated across subjects (subject tenacity), it measures the similarity of different patterns produced across individuals by stimulating the same brain region. A cognitive system's subject tenacity is then the average subject tenacity across regions within the system. When tenacity is calculated across brain regions (region tenacity), it measures the similarity of different patterns produced by stimulation of different brain regions within a given cognitive system in a single individual. This value is then averaged over all individuals. Consequently, a high value of tenacity indicates a high similarity between synchronization patterns produced by stimulation across individuals (subject tenacity) or brain regions (region tenacity). In Fig. 6, we plot each cognitive system based on its score for subject and region tenacity. Additionally, we group cognitive systems by applying a clustering algorithm (see Methods), and the color and shape of a region's icon reflects its group assignment. We identify four distinct groups of systems that are characteris- tically different from each other in terms of their location in the tenacity space and also in their cognitive roles. To better delineate the four groups, we partition the tenacity space based on the level of subject and node tenacity. We define two levels of subject tenacity: variable and stable, while the node tenacity is 12 Figure 6: Classification of cognitive systems based on pattern tenacity. To estimate the consistency of emergent patterns of synchronization within cognitive systems, we constructed a measure called tenacity that estimates the similarity between a set of patterns. Within a cognitive system, we calculate tenacity across two dimensions: patterns that are produced after stimulating each region across subjects (subject tenacity), and patterns that are produced after stimulating different regions of the system within each subject (region tenacity). In the parameter space constructed along these dimensions, we can cluster cognitive systems into 4 groups that suggest a 2x3 partitioning of the tenacity space. This partitioning allows us to dissociate subject-specific and region-specific variability in observed patterns. 13 partitioned into three levels: diverse, flexible, and consistent. Across the subject tenacity dimension, we observe four individually-variable systems that demonstrate the largest variability in patterns. These systems include the frontoparietal and cingulo-opercular, the cognitive control systems that have previously been shown to have large individual variability [47], and ventral temporal association and motor and somatosensory systems that show learning-dependent changes [48, 45, 49]. The remaining five systems are classified as individually-stable with high subject tenacity scores. The default mode, subcortical, and attention systems have previously been found preserved across individuals as well as across species [50], whereas the auditory and visual systems support fundamental perceptual processing [51, 52]. Across the region tenacity dimension, we observe three levels of tenacity. Congitively-consistent systems include the attention, subcortical, and visual systems, and stimulation to regions within these systems give rise to similar patterns of synchrony and de-synchrony. The auditory emerges as the sole cognitively-diverse system. This reflects the starkly different patterns that arise after stimulation of the regions within this system. For example, stimulation to the superior temporal region results in high synchrony (both coherent and chimera) states, while stimulation to the transverse temporal region leads to a metastable state (see Fig. S6). Thus, the different local connectivity patterns of regions within this system produce immense diversity in the resulting synchrony patterns upon their stimulation. The remaining five cognitive systems are classified as cognitively-flexible, indicating that stimulation of regions within these systems produced variable patterns of synchrony. Overall, the tenacity scores both confirm and extend our knowledge of brain structure-function relation- ships. The variability in subject tenacity among the systems reflects known differences in system stability between individuals, and confirms that variability in chimera patterns captures these coarse differences among the systems. On the other hand, the spread of region tenacity scores captures the diversity in the functional roles that the regions within a system serve across diverse cognitive tasks. Cognitively-consistent systems can largely be involved in the core sensory processing and associative learning, whereas the variabil- ity of patterns within cognitively-flexible systems may enable them to serve diverse cognitive roles, relying on stimulation of each constituent as a means to synchronize and integrate with different cognitive systems to support particular cognitive demand or task-relevant processing. Discussion Using a novel, chimera-based framework, we explored the dynamical states that emerge across large-scale cognitive systems following the spread of a targeted regional stimulation. We identified three distinct dy- namical states – coherent, chimera, and metastable – that arise as a function of the structural connectivity of the stimulated regions. A core result across all analyses is the variety in frequency and distribution of the observed dynamical states. Chimera states are the most pervasive state to emerge following regional stimu- lation. This likely reflects the foundational role that partial synchrony serves in large-scale brain function to enable the intricate balance between segregated and integrated neural processing. Furthermore, the diver- sity in these patterns captured both subject-specific and region-specific variability in structural connectivity. Based on its sensitivity to these different sources of variability, our novel chimera-based framework shows immense promise to better understand individual differences, relating patterns to performance, as well as system constraints that underlie how to drive the brain to different task-relevant states. 14 Prevalence of chimera states across cognitive systems enables segregation and integration in brain dynamics The brain is a complex dynamical system that must integrate information across spatially-distributed, seg- regated regions that serve specialized functional roles [53]. Neuroscience research therefore attempts to understand how the brain creates selective synchrony across subsets of task-relevant regions to enable rapid and adaptive cognitive processing [54, 55]. Recently, network neuroscience approaches have identified sets of brain regions that form cognitive systems during rest and task states [56]. By studying the interactions of these cognitive systems, functional analyses have identified the importance of (a) integrated states where the connections are stronger between cognitive systems[12] and (b) segregated states where connections are weaker between cognitive systems and are likely to be stronger within. The relative level of func- tional integration vs segregation of cognitive systems has important consequences for cognitive performance. Highly segregated systems enable efficient computations in local, functionally-specialized brain regions, while strongly integrated systems provide rapid consolidation of information across systems necessary for coordi- nated, cohesive performance of complex tasks [12, 13, 10]. These critical brain states are captured in the chimera framework as metastable (segregated) states and coherent (integrated) states, and perhaps the most critically for brain function, the chimera state that describes partial synchrony across subsets of cognitive systems. All nine systems give rise to a chimera state following stimulation, suggesting that all cognitive systems can drive the brain to partial synchrony in support of their functional roles in cognition. Our results augment a burgeoning literature on what brain dynamics support rapid shifts between more segregated or integrated brain states. Previous work has found that functionally segregated states tend to involve shorter, local connections [57], while integration largely relies on the global influence of subcortical regions and cortical hubs that have many diverse connections to other brain regions [58]. Collectively, our results demonstrate that our novel chimera-framework can investigate critical cognitive states where a balance between integration and segregation is required for adaptive cognition. Chimera framework reveals subject-specific and region-specific variability in brain connectivity through the analyses of emergent dynamical states We found that a coherent state is likely to be produced by the stimulation of regions in the medial default mode and subcortical systems. This reflects the propensity of these systems to contain regional hubs, and the prevalent emergence of a coherent state reflects their functional roles to bridge spatially disperse regions and facilitate global brain communication. Interestingly, these systems also have high subject tenacity, indicating robust occurrence across the 30 individuals. This could reflect that the subcortical and medial default mode systems provide a fundamental, constant pillar of brain organization, which when disrupted, could lead to impairments in global brain function. Previous research has shown that network hubs are often found to be impacted in neurological disorders such as schizophrenia [59] and Alzheimer's disease [60]. These disorders are associated with network-wide deficits in brain function [26], which is consistent with our finding that cognitive systems that produce coherent states also contain network hubs. Conversely, metastable states are preferentially produced by four systems with more sparse structural connectivity: two systems associated with cognitive control (cingulo-opercular and frontoparietal systems) and two systems associated with intricate sensory, object, and language representations (auditory and ventral temporal association systems). These systems all have functional roles that frequently require working in seclusion from other specialized processing in the brain. Interestingly, the three systems that are the most unlikely be synchronized upon stimulation are also the ones that are most likely to produce a metastable state: the auditory, cingulo-opercular, and frontoparietal systems. The cingulo-opercular and forntoparietal systems are associated with cognitive control, and they are proposed as complementary systems that are 15 specialized for guiding successful task performance at different timescales: the cingulo-opercular system for maintaining task-related goals across trials, and the frontoparietal systems for trial-by-trial control [61]. These functional roles may often need to occur concurrently, and their production of metastable states could indicate that these systems can work in seclusion without co-activating a large part of the brain, an attribute that facilitates parallel processing. While our chimera framework revealed stable features of brain architecture, it also captured cognitive systems where between subject variability leads to variety in frequency and type of synchronization patterns: cingulo-opercular, frontoparietal, ventral temporal association, and motor systems. These four systems with low subject tenacity are associated with higher cognitive functions where an individuals experience and knowledge are likely captured by variability in their structural connectivity [62]. Our results demonstrated that frontoparietal and cingulo-opercular systems exhibit the strongest individual variability, and this mirrors recent results that showed cognitive control systems have weaker within-subject variability and greater between-subject variability relative to sensory processing systems [47]. Our results also demonstrated that the ventral temporal association and motor/somatosensory systems show an especially high number of prevalent patterns with no single dominant pattern, and this may reflect their roles in learning and development-related changes [49, 63]. Methodological Considerations and Future Directions Our model is only sensitive to functional relationships that are induced through structural connections, so the observed dynamical states and patterns are only constrained by the anatomical structure of the network. In reality, neuronal activity patterns that are observed in brain using different functional measurement techniques, such as fMRI, EEG, MEG, and PET, are a result of a complex neurophysiological activity that develops on top of the structural connectivity infrastructure. Thus, the actual patterns of brain activity that are observed across functional modalities may come from the simultaneous activation of different brain regions via multiple input sources and therefore might differ from the patterns observed in our in silico experiments. Here, the emergence of a coherent pattern would imply that a node can, in principle, communicate with all of the spatially distributed regions within the brain; however, the actual nodes that it communicates with may vary between different tasks according to the specific cognitive demands of the task. Likewise, the emergence of a metastable state in our model may not reflect total de-synchrony, but just synchronization at a smaller population level that our framework simulates, requiring future models to study dynamical states at a finer spatial resolution. Despite these limitations, our approach is sensitive to variability in region-specific and subject-specific brain connectivity, and it can be used to answer fundamental questions concerning the cognitive organization of the human brain necessary for quantifying meaningful individual differences in brain architecture, supporting individualized medicine, performance enhancement technologies, and personalized stimulation protocols for treatment and/or augmentation. Conclusion By employing the cognitive system framework [37], our novel chimera-based approach keeps the computa- tional complexity of the analysis to a minimum while confirming the existence of chimera states in the large- scale cognitive systems in the human brain, as predicted from low-level, small scale models [21, 22, 32, 33]. The partial synchrony observed in a chimera state has a natural link to the well-documented role of functional segregation and integration of cognitive systems thought to support cognition [54, 14], and the approach cap- tures robust system differences for those that are largely stable across people as well as those that capture individual training and expertise. Thus, our approach provides a rich opportunity to study how dynamical states give rise to variability in cognitive performance, providing the first link between how chimera states may subserve human behavior. 16 Methods Human diffusion spectrum imaging data acquisition and pre-processing. Diffusion MRI analysis was performed on the 30 individual participant scans previously reported elsewhere [64]. Twenty male and ten female subjects were recruited from Pittsburgh and the Army Research Laboratory in Aberdeen, Maryland. All subjects were neurologically healthy, with no history of either head trauma or neurological or psychiatric illness. Subject ages ranged from 21 to 45 years of age at the time of scanning (mean age of 31 years) and four were left-handed (2 male, 2 female). All participants signed an informed consent approved by Carnegie Mellon University and conforming with the Declaration of Helsinki and were financially remunerated for their participation. Macroscopic white matter pathways were imaged using a Diffusion Spectrum Imaging (DSI) aquisition sequence on a Siemens Verio 3T MRI system located at the Scientific Imaging & Brain Research Center (SIBR) at Carnegie Mellon University using a 32-channel head coil. A total of 257-direction were sampled using a twice-refocused spin-echo sequence (51 slices, TR = 9.916s, TE = 157ms, 2.4 x 2.4 x 2.4mm voxels, 231 x 231mm FoV, and b-max = 5,000s/mm2). Diffusion data were reconstructed using q-space diffeomorphic reconstruction (QSDR;[65]) with a diffusion sampling length ratio of 1.25 and a 2mm output resolution. Construction of individual structural brain networks. Whole-brain structural connectivity matrices were constructed for each subject using a bootstrapping ap- proach. To minimize the impact of bias in the tractography parameter scheme on streamline generation, whole-brain fiber tractography [66] was performed 100 times for each participant, generating 250,000 stream- lines per iteration. Across the 100 iterations, values were randomly sampled for QA-based fiber termination thresholds (0.01-0.10), turning angle thresholds (40 -80 ), and smoothing (50%-80%), while constant values were used for step size (1mm) and min/max fiber length thresholds (10mm/400mm). On each iteration a binary connectivity matrix was generated, where an edge between two regions of interest was considered present if 5% or more of streamlines generated were found to connect them. The probability of observing a connection was estimated by calculating the frequency of detecting an edge across all 100 iterations. The region of interest were determined using cortical components of the Desikan-Killiany atlas and subcorti- cal components of the Harvard-Oxford Subcortical atlas. In the resulting weighted matrices, connection strengths were normalized by the sum of the regional brain volumes, and these normalized matrices were used as the structural representations of individual brains. All analysis was performed using DSI Studio (http://dsi-studio.labsolver.org/) and Matlab (MathWorks, Inc.; Natick, MA, USA). Data-driven network model of brain dynamics. In our data-driven network model, regional brain dynamics are given by Wilson-Cowan oscillators [39, 36, 40]. In this biologically motivated neural mass model, the fraction of excitatory and inhibitory neurons active at time t in the ith brain region are denoted by Ei(t) and Ii(t) respectively, and their temporal dynamics are given by: τ dEi dt = −Ei(t) + (SEm − Ei(t))SE c1Ei(t) − c2Ii(t) + c5 AijEj(t − τ ij d ) + Pi(t) + σwi(t), (1) τ dIi dt = −Ii(t) + (SIm − Ii(t))SI c3Ei(t) − c4Ii(t) + c6 AijIj(t − τ ij d ) + σvi(t), (2) (cid:16) (cid:16) (cid:88) (cid:88) j (cid:17) (cid:17) j 17 where SE,I (x) = 1 1 + e(−aE,I (x−θE,I ) − 1 1 + eaE,I θE,I . (3) Aij is an element of the subject-specific coupling matrix A whose value is the connection strength between brain regions i and j as determined from diffusion spectrum imaging as described above. The global strength of coupling between brain regions is tuned by excitatory and inhibitory coupling parameters c5 and c6 re- spectively. In this case c6 = c5/4. Pi(t) represents the external stimulation to excitatory state activity and was used to perform computational stimulation experiments. The parameter τ ij d represents the communi- cation delay between regions i and j. If the spatial distance between regions i and j is dij, τ ij d = dij/td, where td = 10m/s is the signal transmission velocity. We added noise as an input to the system through the parameters wi(t) and vi(t) which are derived from a normal distribution with σ = 10−5. Other constants in the model are biologically derived: c1 = 16, c2 = 12, c3 = 15, c4 = 3, aE = 1.3, aI = 2, θE = 4, θI = 3.7, τ = 8 as described in references [39, 36, 40]. To numerically simulate the dynamics of the system, we used a second order Runge Kutta method with step size 0.1 with initial conditions (Ei(0), Ii(0) = 0.1, 0.1). Targeted stimulation. The model was optimized for each individual to allow a regime of maximum dynamical sensitivity. This was done by choosing a global coupling parameter, c5, such that the system was just below the critical transition point to the excited state (see Fig. S1). Regional stimulation was achieved by applying a constant external input Pi = 1.15 to a single region and perturbing its dynamics (Fig. S2). As the dynamics evolve, the stimulation spreads within the brain through the network connectivity of the stimulated node. Cognitive systems. We assigned each brain region to one of 9 cognitive systems: attention, auditory, cingulo-opercular, fron- toparietal, medial default mode, motor and somatosensory, subcortical, ventral temporal association and visual. This node assignment is described in Table S1 and is similar to the one used by Muldoon et al. [36]. The distribution of brain regions within cognitive systems is shown in Fig. S3. Calculation of synchronization within cognitively-informed framework. We used the standard order parameter ρ to estimate the extent of synchronization after a targeted regional stimulation within the brain networks. This measure was proposed by Kuramoto for the estimation of coherence in a population of Kuramoto phase oscillators [42]. In this case, the instantaneous order parameter at a given time t it is defined as ρN (t)eiΦ(t) = 1 N eiφj (t), N(cid:88) j=1 where φj is the phase of the jth oscillator at time t and is given by φj(t) = tan−1 Ij(t) Ej(t) . Here, N = 76 is the number of oscillators in the system. In order to estimate the global synchronization in the system, one needs to average the instantaneous order parameter for a sufficiently long period of time, ρN =< ρN (t) >t . (6) 18 (4) (5) We used simulated activity over 1 s to estimate the average order parameter. Within our cognitively-informed framework, we measured the synchronization between all pairs of cog- nitive systems following a regional stimulation. This was done by calculating an order parameter for the combined oscillator population of a pair of cognitive systems. For cognitive systems ci and cj, this order parameter is given by where ρci,cj (t)eiΘ(t) = 1 Nci + Ncj k∈(ci∪cj ) ρci,cj =< ρci,cj (t) >t, (cid:88) eiφk(t). (7) (8) Here, Nci and Ncj represent the number of oscillators (brain regions or nodes) within cognitive systems ci and cj respectively. This analysis resulted in synchronization matrices, as shown in Fig. 2a, whose entries represent the extent of synchronization between cognitive systems. These matrices were used in order to identify the dynamical cognitive state that emerged as a result of regional stimulation. Chimera-index. We calculated the chimera-index (C) as described in [43, 34] as a measure of the normalized average variation in order parameter within cognitive systems over time. For ci ∈ [c1, c2, ..., cM ] (M = 9 is the total number of systems), where σch(t) = 1 M − 1 (ρci (t)− < ρc(t) >M )2. < σch(t) >t , C = CM ax M(cid:88) i=1 (9) (10) In this case, CM ax = 5/36 is a normalization factor and represents the maximum value of variability in the order parameter in an ideal chimera state where the network organizes such that the half of its population is completely synchronized and half is completely desynchronized [43]. The instantaneous quantity < ρc(t) >M measures the synchronization of cognitive systems, averaged over all systems at a given time t. Robust detection of the patterns of synchronization. In order to robustly identify emergent cognitive patterns, we first obtained a binarized synchronization matrix (s) such that, sij = 1 if systems i and j are identified synchronized, and sij = 0 otherwise. We defined two cognitive systems i and j to be synchronized if ρci,cj ≥ ρT h, where ρT h represents a synchronization threshold. For the results discussed in the main text we used ρT h = 0.8 [43] (as indicated in Fig. 2a). In principle, one can directly use such binarized synchronization matrices in order to classify the emergent states and patterns. However, we performed community detection on these binarized matrices. This method clusters the group of synchronized systems into a single community whereas desynchronized systems remain separate communities. In case of a coherent state, we observed only one community, and in case of a metastable state we observed nine separate communities, each representing a cognitive system. For chimera states, communities with different distributions of cognitive systems emerged. Thus, applying the community detection algorithm not only allowed us to robustly classify the emergent dynamical states, but also let us separate various spatially distributed patterns of chimera states. Community detection was performed using 19 modularity maximization through the generalized Louvain algorithm [67]. For community detection, the value of the resolution parameter was varied between 0.8 to 0.95 and a consensus was run to determine the community structure [68]. Pattern tenacity. A pattern in our analysis describes if the given cognitive system falls into the synchronized population or remains desynchronized (Fig. 5). In order to calculate similarity between patterns, we defined tenacity for a set of observed patterns as follows p(cid:88) (cid:0) 1 M(cid:88) (cid:1), δc i,j T = 1 p(p − 1) i,j=1 M c=1 (11) where p is the number of patterns in the set. δc i,j = 1 if the cognitive system c falls into the same state of either synchrony or desynchrony in patterns i and j, and = 0 otherwise. We calculated the tenacity of cognitive systems in two dimensions: across subjects for a given brain region within the system (subject tenacity), and across regions of the system in a given subject (region tenacity). For subject tenacity of a given cognitive system, p constitutes the patterns that are produced by a given node for all the subjects and equals the number of subjects i.e. 30. For region tenacity of a given cognitive system, p constitutes the patterns that all the nodes for a given cognitive system produce for a given subject, and the value of p varies between systems. Thus for each cognitive system we obtained two distributions of tenacity, one for each subject and region tenacity. Clustering of cognitive systems using tenacity features. In the subject-node tenacity parameter space, we grouped cognitive systems into clusters using the k-means algorithm and silhouette analysis. We used k = 3, 4, 5, and 6 and identified the stable clustering that maximizes similarity within clusters and dissimilarity across clusters. For k = 4 we observed an optimized clustering. The corresponding silhouette plot is shown in Fig. S7. Rendering of brain images. BrainNet Viewer was used to perform spatial mapping onto brain images [69]. References [1] Winfree, A. T. Biological rhythms and the behavior of populations of coupled oscillators. J. Theoret. Biol 16, 15–42 (1967). [2] Glass, L. Synchronization and rhythmic processes in physiology. Nature 410, 277–284 (2001). [3] Jia, T., Wang, D. & Szymanski, B. K. Quantifying patterns of research-interest evolution. Nature Human Behaviour 1, 0078 (2017). [4] Strogatz, S. H., Abrams, D. M., McRobie, A., Eckhardt, B. & Ott, E. Crowd synchrony on the millennium bridge. Nature 438, 43–44 (2005). [5] Bertolero, M. A., Yeo, B. T. T. & D'Esposito, M. The modular and integrative functional architecture of the human brain. Proceedings of the National Academy of Sciences 112, E6798–E6807 (2015). 20 [6] Motter, A. E. Nonlinear dynamics: Spontaneous synchrony breaking. Nature Physics 6, 164–165 (2010). [7] Pecora, L. M., Sorrentino, F., Hagerstrom, A. M., Murphy, T. E. & Roy, R. Cluster synchronization and isolated desynchronization in complex networks with symmetries. Nature Communications 5, 4079 (2014). [8] Gray, C. M. Synchronous oscillations in neuronal systems: Mechanisms and functions. Journal of Computational Neuroscience 1, 11–38 (1994). [9] Lachaux, J. P., Rodriguez, E., Martinerie, J. & Varela, F. J. Measuring phase synchrony in brain signals. Hum. Brain Mapp. 8, 194–208 (1999). [10] Bullmore, E. & Sporns, O. Complex brain networks: graph theoretical analysis of structural and functional systems. Nature Reviews Neuroscience 186–198 (2009). [11] Bassett, D. S. & Sporns, O. Network neuroscience. Nature Neuroscience 20, 353–364 (2017). [12] Shine, J. M. et al. The Dynamics of Functional Brain Networks: Integrated Network States during Cognitive Task Performance. Neuron 92, 544–554 (2016). [13] Vatansever, D., Menon, D. K. & Stamatakis, E. A. Default mode contributions to automated information processing. Proceedings of the National Academy of Sciences of the United States of America 114, 12821–12826 (2017). [14] Deco, G., Tononi, G., Boly, M. & Kringelbach, M. L. Rethinking segregation and integration: Contri- butions of whole-brain modelling. Nature Reviews Neuroscience 16, 430–439 (2015). [15] Kuramoto, Y. & Battogtokh, D. Coexistence of coherence and incoherence in nonlocally coupled phase oscillators. Nonlinear Phenomena in Complex Systems 5, 380–385 (2002). [16] Abrams, D. M. & Strogatz, S. H. Chimera states for coupled oscillators. Phys. Rev. Lett. 93, 174102 (2004). [17] Omelchenko, O. E., Maistrenko, Y. L. & Tass, P. A. Chimera states: The natural link between coherence and incoherence. Phys. Rev. Lett. 100, 044105 (2008). [18] Laing, C. R. The dynamics of chimera states in heterogeneous kuramoto networks. Physica D: Nonlinear Phenomena 238, 1569 – 1588 (2009). [19] Hagerstrom, A. M. et al. Experimental observation of chimeras in coupled-map lattices. Nature Phys. 8, 658–661 (2012). [20] Tinsley, M. R., Nkomo, S. & Showalter, K. Chimera and phase-cluster states in populations of coupled chemical oscillators. Nature Physics 8, 662–665 (2012). [21] Omelchenko, I., Omel'chenko, O. E., Hovel, P. & Scholl, E. When nonlocal coupling between oscillators becomes stronger: Patched synchrony or multichimera states. Phys. Rev. Lett. 110, 224101 (2013). [22] Vllings, A., Hizanidis, J., Omelchenko, I. & Hovel, P. Clustered chimera states in systems of type-i excitability. New Journal of Physics 16, 123039 (2014). [23] Xie, J., Knobloch, E. & Kao, H.-C. Multicluster and traveling chimera states in nonlocal phase-coupled oscillators. Phys. Rev. E 90, 022919 (2014). 21 [24] Zakharova, A., Kapeller, M. & Scholl, E. Amplitude chimeras and chimera death in dynamical networks. Journal of Physics: Conference Series 727, 012018 (2016). [25] Killgore, W. D. S. Effects of sleep deprivation on cognition. Progress in Brain Research 185, 105129 (2010). [26] Menon, V. Large-scale brain networks and psychopathology: A unifying triple network model. Trends in Cognitive Sciences 15, 483–506 (2011). [27] Fornito, A., Zalesky, A. & Breakspear, M. The connectomics of brain disorders. Nature Reviews Neuroscience 16, 159–172 (2015). [28] Panaggio, M. J. & Abrams, D. M. Chimera states: Coexistence of coherence and incoherence in networks of coupled oscillators. Nonlinearity 28, R67 (2015). [29] Andrzejak, R. G., Rummel, C., Mormann, F. & Schindler, K. All together now: Analogies between chimera state collapses and epileptic seizures. Scientific reports 6, 23000 (2016). [30] Tognoli, E. & Kelso, J. A. S. The metastable brain. Neuron 81, 35–48 (2014). [31] Ashwin, P., Coombes, S. & Nicks, R. Mathematical frameworks for oscillatory network dynamics in neuroscience. The Journal of Mathematical Neuroscience 6, 2 (2016). [32] Hizanidis, J., Kanas, V. G., Bezerianos, A. & Bountis, T. Chimera states in networks of nonlocally coupled hindmarshrose neuron models. International Journal of Bifurcation and Chaos 24, 1450030 (2014). [33] Bera, B. K., Ghosh, D. & Lakshmanan, M. Chimera states in bursting neurons. Phys. Rev. E 93, 012205 (2016). [34] Hizanidis, J., Kouvaris, N. E., Gorka, Z.-L., D´ıaz-Guilera, A. & Antonopoulos, C. G. Chimera-like States in Modular Neural Networks. Scientific Reports 6, 19845 (2016). 1510.00286. [35] Santos, M. et al. Chimera-like states in a neuronal network model of the cat brain. Chaos, Solitons and Fractals 101, 86 – 91 (2017). [36] Muldoon, S. F. et al. Stimulation-based control of dynamic brain networks. PLoS Comput Biol 12, e1005076 (2016). [37] Power, J. D. et al. Functional network organization of the human brain. Neuron 72, 665–678 (2011). [38] Yeh, F.-C. et al. Quantifying Differences and Similarities in Whole-Brain White Matter Architecture Using Local Connectome Fingerprints. PLoS Comput Biol 12 (2016). [39] Wilson, H. R. & Cowan, J. D. Excitatory and inhibitory interactions in localized populations of model neurons. Biophys J 12, 1–24 (1972). [40] Bansal, K., Medaglia, J. D., Bassett, D. S., Vettel, J. M. & Muldoon, S. F. Data-driven brain network models predict individual variability in behavior Preprint. Available from: arXiv:1802.08747v1. Cited 27 April 2018. (2018). [41] Bansal, K., Nakuci, J. & Muldoon, S. F. Personalized brain network models for assessing structure- function relationships. Current Opinion in Neurobiology 52, 42 – 47 (2018). 22 [42] Kuramoto, Y. Self-entrainment of a population of coupled non-linear oscillators. In Araki, H. (ed.) Lecture Notes in Physics, International Symposium on Mathematical Problems in Theoretical Physics, 420 (Springer-Verlag, New York, 1975). [43] Shanahan, M. Metastable chimera states in community-structured oscillator networks. Chaos: An Interdisciplinary Journal of Nonlinear Science 20, 013108 (2010). [44] Dehaene, S., Cohen, L., Morais, J. & Kolinsky, R. Illiterate to literate: behavioural and cerebral changes induced by reading acquisition. Nature Reviews Neuroscience 16, 234–244 (2015). [45] Sheinberg, D. L. & Logothetis, N. K. Noticing familiar objects in real world scenes: the role of temporal cortical neurons in natural vision. J Neurosci 21, 134050 (2001). [46] Grill-Spector, K., Kourtzi, Z. & Kanwisher, N. The lateral occipital complex and its role in object recognition. Vision Res 41, 140922 (2001). [47] Gratton, C. et al. Functional Brain Networks Are Dominated by Stable Group and Individual Factors, Not Cognitive or Daily Variation. Neuron 98, 439–452.e5 (2018). [48] Muraskin, J. et al. Fusing multiple neuroimaging modalities to assess group differences in perception action coupling. Proceedings of the IEEE 105, 83100 (2017). [49] Scholz, J., Klein, M. C., Behrens, T. E. J. & Johansen-Berg, H. Training induces changes in white-matter architecture. Nature Neuroscience 12, 13701371 (2009). [50] Margulies, D. S. et al. Precuneus shares intrinsic functional architecture in humans and monkeys. Proceedings of the National Academy of Sciences 106, 20069–20074 (2009). [51] Kaskan, P. M. et al. Peripheral variability and central constancy in mammalian visual system evolution. Proceedings. Biological sciences 272, 91–100 (2005). [52] Stevens, C. F. An evolutionary scaling law for the primate visual system and its basis in cortical function. Nature 411, 193–195 (2001). [53] Alivisatos, A. P. et al. The brain activity map project and the challenge of functional connectomics. Neuron 74, 970–4 (2012). [54] Menon, V. Large-scale brain networks in cognition: emerging methods and principles. Trends Cogn Sci 14, 277–290 (2010). [55] Gollo, L. L., Roberts, J. A. & Cocchi, L. Mapping how local perturbations influence systems-level brain dynamics. NeuroImage 160, 97–112 (2017). [56] Smith, S. M. et al. Correspondence of the brains functional architecture during activation and rest. Proceedings of the National Academy of Sciences 106, 1304013045 (2009). [57] Li´egeois, R. et al. Cerebral functional connectivity periodically (de)synchronizes with anatomical con- straints. Brain Structure and Function 221, 2985–2997 (2016). [58] Shine, J. M., Aburn, M. J., Breakspear, M. & Poldrack, R. A. The modulation of neural gain facilitates a transition between functional segregation and integration in the brain. eLife 7 (2018). [59] Klauser, P. et al. White matter disruptions in schizophrenia are spatially widespread and topologically converge on brain network hubs. Schizophrenia Bulletin 43, 425–435 (2017). 23 [60] Buckner, R. L. et al. Cortical hubs revealed by intrinsic functional connectivity: Mapping, assessment of stability, and relation to alzheimer's disease. Journal of Neuroscience 29, 1860–1873 (2009). [61] Cocchi, L., Zalesky, A., Fornito, A. & Mattingley, J. B. Dynamic cooperation and competition between brain systems during cognitive control. Trends in Cognitive Sciences 17, 493–501 (2013). [62] Powell, M. A., Garcia, J. O., Yeh, F.-C., Vettel, J. M. & Verstynen, T. Local connectome phenotypes predict social, health, and cognitive factors. Network Neuroscience 2, 86–105 (2018). [63] Kahn, A. E. et al. Structural pathways supporting swift acquisition of new visuomotor skills. Cerebral Cortex 27, 173–184 (2017). [64] Verstynen, T. D. The organization and dynamics of corticostriatal pathways link the medial orbitofrontal cortex to future behavioral responses. Journal of neurophysiology 112, 2457–2469 (2014). [65] Yeh, F.-C. & Tseng, W.-Y. I. Ntu-90: a high angular resolution brain atlas constructed by q-space diffeomorphic reconstruction. Neuroimage 58, 91–99 (2011). [66] Yeh, F.-C., Verstynen, T. D., Wang, Y., Fern´andez-Miranda, J. C. & Tseng, W.-Y. I. Deterministic diffusion fiber tracking improved by quantitative anisotropy. PloS one 8, e80713 (2013). [67] Mucha, P. J., Richardson, T., Macon, K., Porter, M. A. & Onnela, J.-P. Community structure in time-dependent, multiscale, and multiplex networks. Science 328, 876–878 (2010). [68] Bassett, D. S. et al. Robust detection of dynamic community structure in networks. Chaos: An Interdisciplinary Journal of Nonlinear Science 23, 013142 (2013). [69] Xia, M., Wang, J. & He, Y. BrainNet Viewer: A Network Visualization Tool for Human Brain Con- nectomics. PLoS ONE 8, e68910 (2013). Acknowledgments This work is supported by the Army Research Laboratory through contract # W911NF-10-2-0022 and W911NF-16-2-0158 from the U.S. Army research office. The authors acknowledge Gregory Lieberman for the pre-processing of the brain anatomical connectivity data, and Rajarshi Roy and Joseph D. Hart for a useful discussion on chimera states. The content is solely the responsibility of the authors and does not necessarily represent the official views of the funding agency. Author contributions KB conceived the idea; KB, JMV, and SFM designed the research; TV and JMV collected the data; SFM developed the model; KB implemented the model and led the analysis; all authors contributed analysis ideas and wrote the paper. 24 Figure S1: Optimizing personalized brain network models. In our model, coupling between different brain regions can be tuned by the global coupling parameter,c5. We optimized this parameter for each individual to place the brain in a dynamical state that is maximally sensitive to a perturbation (stimulation). To find this optimal value for each individual, we varied c5 between 100 and 1500 in steps of 10 and measured average excitatory activity (Ei) of each brain region (i) for a time period of 1 s. Each simulation was started with (Ei , Ii) = (0.1, 0.1), and before starting the measurement, we let the system evolve for 1 s in order to eliminate initial transients. (a) As we varied the value of c5, we observed a sudden transition in systems behavior to an excited state. (b) The value of c5 at which the transition is observed typically varies between individuals. The transition value signifies the ease with which a given brain can be excited. In order to apply a computational stimulation within the personalized model, we fixed the value of c5 just below its transition value. 25 Figure S2: Targeted regional stimulation. (a) To stimulate a particular brain region within the brain network model of a given individual, we applied a constant external input, Pi = 1.15. In our model, as we switch on the external input, the stimulated brain region changes its dynamics from a (b) stable fixed point to (c) a limit cycle (oscillatory motion). The effect of this stimulation on the other regions within the brain is measured through the resulting patterns of synchronization as discussed in the main text. 26 Figure S3: Distribution of brain volume within cognitive systems. Spatial mappings showing the distribution of the nine cognitive systems within the brain. The identification of regions is given in Table S1. L and R denote the left and right hemisphere respectively. 27 Figure S4: Effect of changing synchronization threshold on the distribution of states. (a)-(d) Separation of three different states in the parameter space of global synchrony and chimera-index for different synchro- nization threshold values (ρT h). Using different values of ρT h in our analysis, we obtain results that are qualitatively similar to the results discussed in Fig. 2b. We observe three different states i.e. coherent, chimera and metastable states, for all four threshold values. These states can be clearly separated in the parameter space of global synchrony and chimera-index. However, a higher threshold value (e.g. 0.9), which signifies strictness in identifying a population to be synchronized, allows a lower number of states to be clas- sified as a coherent state, as opposed to a lower threshold value (e.g. 0.7), which is more relaxed. Conversely, number of observed metastable states increases with increasing threshold values. 28 Figure S5: Likelihood of the emergence of dynamical states across cognitive systems. Contribution of different cognitive systems towards producing a given cognitive states is depicted for (a) ρT h = 0.75 and (b) ρT h = 0.85. These figures corroborate our findings presented in Fig. 4. We observe that a coherent state dominantly originates when the regions in subcortical and medial default mode systems are stimulated. A metastable state is likely to originate upon stimulation of the regions within cingulo-opercular, frontoparietal, auditory and ventral temporal association systems. Chimera states do not show the dominance of any particular cognitive system, regions from different cognitive systems are relatively equally likely to produce of chimera state upon stimulation. 29 30 31 Figure S6: Normalized contribution of brain regions to the prevalent patterns of synchronization. In Fig. 5 of the main text we show the prevalent patterns of synchronization that we observed after systematic regional stimulation within the brain across 30 individuals. Here, we show the brain regions (nodes) that contributed in producing these patterns within each cognitive system. Each sub-figure represents a cognitive system and the first image depicts all the brain regions within the system. In the following images, we show only the brain regions (nodes) that produced the corresponding patterns in Fig. 5. The size of the node represents the normalized occurrence for the given pattern. In panel (b), in order to help the reader differentiate superior and transverse regions, we mark them with S and T respectively. From this figure it can be clearly observed that different nodes within the same cognitive system can produce different patterns. Normalized occurrences for different patterns also vary across patterns and across cognitive systems. 32 Figure S7: Clustering of cognitive systems using patterns tenacity. As described in the main text, in the parameter space of subject and node tenacity we grouped cognitive systems into clusters using the k-means algorithm and silhouette analysis. For k = 4 we observed an optimized clustering (Fig. 6). Here we show the corresponding silhouette plot using squared Euclidean distance of the data points from the centroid of the cluster to which they were assigned. 33 S.N. Name of the brain region Cognitive assignment 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 Lateral orbitofrontal Superior parietal Superior temporal Transverse temporal Caudal anterior cingulate Pars opercularis Pars orbitalis Rostral anterior cingulate Rostral middle frontal Supramarginal Caudal middle frontal Inferior parietal Medial orbitofrontal Pars triangularis Insula Isthmus cingulate Posterior cingulate Precuneus Superior frontal Att Att Aud Aud CP CP CP CP CP CP FP FP FP FP FP mDM mDM mDM mDM Paracentral Postcentral Precentral Thalamus Caudate Putamen Pallidum S.N. Name of the brain region Cognitive assignment 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 MS MS MS SC SC SC SC SC SC SC VT VT VT VT VT V V V V Inferior temporal Middle temporal Parahippocampal Amygdala Accumbens Entorhinal Fusiform Hippocampus Cuneus Lateral occipital Lingual Pericalcarine Table S1: Assignment of brain regions to cognitive systems. Each node in the brain network is assigned to a predefined cognitive system. System assignments are the same for regions in both left and right hemisphere. In Figure S3, we show the distribution of regional brain volume for each cognitive system. 34